Anda di halaman 1dari 79

LECTURE NOTES ON

Physics 18324304
MODERN PHYSICS
Tjipto Prastowo, Ph.D and Endah Rahmawati, M.Si
Department of Physics
Faculty of Mathematics and Natural Sciences
The State University of Surabaya
September 2011
TO THE STUDENT WE LOVE
Lecture notes on Modern Physics contain supporting materials for a series of lectures given
to the second year students of an international class program in the Department of Physics,
Faculty of Mathematics and Natural Sciences, The State University of Surabaya once a week.
These notes are simply intended to focus on two major divisions in Modern Physics: Theory of
Relativity and Theory of Quantum. Thus, these notes are written to provide simple but clear,
mathematically uncomplicated explanations of physical concepts of the two divisions. Each
chapter in these notes is accompanied with some exercises suitable for students homework
assignments. In order to master materials covered in these notes, you need not just knowledge
but skill. This can only be obtained through continual practices. You may obtain a supercial
knowledge by listening to the lectures, but you cannot reach the skill expected by that way.
It is common to come across student conversation, like this I understand it but I cant do
the problem ! This student feels uncomfortable with some problems although they look so
easy when teachers describe them in class.
The above example shows lack of practice and hence lack of skill required in this course.
Our dearest students, please always study with pencil and paper at hand. You will nd that
the more able you are to choose eective methods of solving problems the easier it will be for
you to master new materials. This costs you nothing but practice, practice and again practice.
Please do remember that the best way to learn to solve problems is to solve them.
We eventually welcome good comments on the content of this course from all readers for
further improvement of these notes as the availability of the notes is important to improve
the quality of learning and teaching processes, particularly in the course of Modern Physics.
Hope these notes are useful for all users in the department.
All the best,
Kampus Ketintang 11 September 2011
T jipto Prastowo, Endah Rahmawati
ii
iii
General Guidance
PHYSICS 18324304: MODERN PHYSICS
Pre-requisites: Fundamental Physics (I and II) and Mathematical Physics (I and II)
Lecturers: Tjipto Prastowo, Ph.D and Endah Rahmawati, M.Si
References: Libo, 1980; Beiser, 1988; Gasiorowicz, 1996; Tipler, 1999;
Morin, 2003; Serway et al., 2005; McMahon, 2006; Harris, 2007
Time and Place: Tuesday, 07.00 - 09.30am, D4
Marking Scheme: NA = 20%P + 20%UTS + 30%T + 30%UAS
NA=Final Mark, P=Presence, UTS=Mid-Exam, T=Homework, UAS=Final Exam
Notes:
1. Students are not allowed to join the class for being late (a maximum of 15 minutes from
the starting time is permitted), except for reasonable arguments.
2. Each lecturer contributes an equal proportion of mark to the nal mark.
3. P is possibly reduced to a minimum.
4. UTS = 100% taken from Quiz
5. T = 100% taken from Homework
6. UAS normally contains 4, but possibly has 5 problems.
7. Homework will be distributed to class members and all students are required to hand
the completed assignments in within a given time. Various penalties will be given for
any delay, i.e, 25% discounted mark for a one-day delay and 50% for a two-day delay.
There will be no mark given for those who submit the assignments more than two-day
delays.
8. No additional assignments or examinations after formal exam (both Mid and Final),
except for specied reasons with very limited permission given or medical examination
required.
9. Students are allowed to work with their notes and books in both Mid and Final Exams.
10. Other important issues, if any, will be discussed in the class. Students are strongly
encouraged to be active and well-prepared. If possible, tutorial is available for a further,
detailed description of each topics.
COURSE CONTENTS
1. Chapter One: The Special Theory of Relativity (Weeks 1-5)
overview of modern physics, fundamental aspects of special theory of relativity,
Galilean system vs Lorentz transformation, relativistic four-vector formulation
2. Intermezo: Watching DVD on the Einsteins biography (Week 6)
The DVD provides the remarkable documentary on the life of Albert Einstein,
considered as the greatest scientist since ever after his astonishing contributions to
science and humanity.
3. Chapter Two: The General Theory of Relativity (Weeks 7-9)
Einsteins equivalence principle and its consequences, gravitational time dilation,
gravitational red-shift, the bending of light by the Sun
4. Intermezo: Quiz I on the Relativity Theory (Week 10)
5. Chapter Three: The Black-Body Radiation (Week 11)
classical black-body radiation, Stefan-Boltzmanns law, Wiens displacement law,
Rayleigh-Jeans law, Plancks radiation formula
6. Chapter Four: The Particle Properties of Electromagnetic Radiation (Week 12)
photo-electric eect, Compton eect
7. Chapter Five: The Bohr Model of the Atom (Week 13)
early atomic model, Bohr model of hydrogen atom, Bohrs explanation to his model,
hydrogen spectral lines, Bohrs correspondence principle
8. Chapter Six: The Wave Properties of Sub-Atomic Particles (Week 14)
de Broglie hypothesis and its implications, Davisson-Germer experiment
9. Chapter Seven: The Heisenberg Uncertainty Principle (Week 15)
mathematical basis of the Heisenberg uncertainty principle, its interpretations and
consequences of the uncertainty principle
10. Intermezo: Quiz II on the Quantum Theory (Week 16)
iv
Contents
1 THE SPECIAL THEORY OF RELATIVITY 1
1.1 An Overview of Modern Physics . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Fundamental Aspects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.3 Galilean System versus Lorentz Transformation . . . . . . . . . . . . . . . . . 3
1.4 The Four-Vectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.4.1 The position four-vector . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.4.2 The energy-momentum four-vector . . . . . . . . . . . . . . . . . . . . 7
1.4.3 The rest energy and total relativistic energy . . . . . . . . . . . . . . . 8
1.4.4 The relativistic kinetic energy . . . . . . . . . . . . . . . . . . . . . . . 9
1.4.5 The relativistic Doppler eects . . . . . . . . . . . . . . . . . . . . . . . 10
1.4.6 The wave operator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
1.5 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2 THE GENERAL THEORY OF RELATIVITY 17
2.1 The Equivalence Principle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.2 Consequences of the Equivalence Principle . . . . . . . . . . . . . . . . . . . . 19
2.2.1 Gravitational time dilation . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.2.2 Gravitational red-shift . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.2.3 The bending of light rays by gravity . . . . . . . . . . . . . . . . . . . 22
2.3 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
3 THE BLACK-BODY RADIATION 25
3.1 Thermal Radiation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
3.2 Wiens Displacement Law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
3.3 Rayleigh-Jeans Law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
3.4 Plancks Formula . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.5 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
4 THE PARTICLE PROPERTIES OF ELECTROMAGNETIC RADIATION 31
4.1 Photo-Electric Eect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
4.2 Compton Eect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
4.3 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
5 THE BOHR MODEL OF THE ATOM 39
5.1 The Early Atomic Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
5.2 Bohrs Atomic Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
5.3 Bohrs Explanation to the Model . . . . . . . . . . . . . . . . . . . . . . . . . 42
v
vi CONTENTS
5.4 Hydrogen Spectral Lines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
5.5 Bohrs Correspondence Principle . . . . . . . . . . . . . . . . . . . . . . . . . . 46
5.6 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
6 THE WAVE PROPERTIES OF SUB-ATOMIC PARTICLES 51
6.1 De Broglie Hypothesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
6.2 Implications of the De Broglie Hypothesis . . . . . . . . . . . . . . . . . . . . 54
6.3 The Davisson-Germer Experiment . . . . . . . . . . . . . . . . . . . . . . . . . 56
6.4 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
7 THE HEISENBERG UNCERTAINTY PRINCIPLE 61
7.1 Mathematical Basis for the Uncertainty Principle . . . . . . . . . . . . . . . . 62
7.2 Interpretations of the Uncertainty Principle . . . . . . . . . . . . . . . . . . . 65
7.3 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
A QUIZ on The Relativity Theory 69
B QUIZ on The Quantum Theory 71
Bibliography 73
Chapter 1
THE SPECIAL THEORY OF
RELATIVITY
1.1 An Overview of Modern Physics
Classical physics is responsible for describing many physical phenomena at macroscopic scales
happening around us. This branch of physics is also termed as Newtonian mechanics although
this term is a bit sloppy as it covers not only mechanics but also thermodynamics, waves
and optics, and electricity and magnetism. However, in the early of the twentieth century
revolutionary ideas about the way in which we see our everyday world came about and shook
classical physics to its foundation. It was a time when the existing laws of classical physics
failed to describe the nature of space and time, and of matter and wave at a sophisticated
level.
It was then formally the birth of modern physics, a new paradigm of physics - as opposed
to old views - where space and time and also matter and wave are treated in dierent ways.
Indeed, modern physics relies upon two basic ideas: space and time are not independent and
absolute as they seem to be, and matter may behave as a wave, or vice versa. This new
paradigm can be, for their initially independent paths, divided into two lines of path, that
is theory of relativity and theory of quantum. On one hand, theory of relativity deals with
the dynamics of moving objects at high speeds, approaching the speed of light. This theory is
actually a general form of Newtonian mechanics, in that it agrees well with classical mechanics
for moving bodies at low speeds. On the other hand, theory of quantum discusses the continual
interplay between particles and waves as implied by their dual properties. Quantum physics
puts the deterministic properties of classical physics into the bin, and replaces them with
probabilistic ones. Therefore, there is no way to dene, say an electron, at a specic location
and at a particular time with high condence. What could be here determined is looking for
the probability of nding an electron in a conned area at a certain time with some condence.
1
2 1. THE SPECIAL THEORY OF RELATIVITY
A series of lectures on this course is therefore aimed at understanding relativity theory
and quantum theory in separated ways. In this context, we show that classical mechanics is
merely a special case of the two theories. The speed, not the geometry nor the size, of moving
objects determines whether relativistic theory is required to solve the dynamics of fast-moving
systems. Whilst it is the size, not the speed, of moving objects that calls for quantum theory
to describe the physics of microscopic world.
1.2 Fundamental Aspects
There are in fact two primary themes in Relativity Theory, namely Special Relativity and
General Relativity. While the former does deal with moving objects in inertial frames of
reference, the latter is correct only for non-inertial frames of reference. It is thus important
at this stage to distinguish inertial frames from non-inertial ones. An inertial frame is dened
as a physical coordinate system that is at rest or moving with constant velocity, or in uniform
motion, with respect to a reference, say the Earth for example. In a more delightful denition,
we could say that an inertial frame is the one in which Newtons laws of motion hold. Here,
we focus only on the special theory of relativity that relies upon two basic postulates stated
by Einstein in 1905 as follows :
The principle of relativity, meaning that the laws of physics apply in all inertial
frames.
The universal speed of light, claiming that the speed of light in vacuum is the same for
all inertial observers, regardless of the relative motion between the source and observer.
Various approaches can be taken in formulating consequences of the two relativistic postulates;
dierent approaches taken have their own particular purposes, and no good or bad route to
take as they are equally valid and challenging. The followings are such consequences.
Not a single object in the sluggish world can travel faster than the speed of light in
vacuum, in the sense that the speed of light serves as the upper limit for the speed of
moving objects.
Space and time are actually interdependent quantities, and are not absolute.
Newtonian mechanics needs to be reconsidered when it is applied to moving objects at
high speeds.
There exists dierent ways of thinking about what is called as the consequences of the two
relativistic postulates on one hand, and the eects of those consequences on the dynamics of
1.3. Galilean System versus Lorentz Transformation 3
moving bodies at high speeds on the other hand. We put them together in an ordered section
as discussed here to show the most striking eects of the relativistic consequences below :
loss of simultaneity, describing two events that are simultaneous in one inertial frame
of reference but may not in another inertial frame of reference.
length contraction, derived as a result of the loss of simultaneity.
loss of coincidence in space, describing two events that occur at the same place in
one inertial frame of reference but may not in another inertial frame of reference.
time dilation, derived as a result of the loss of spatial coincidence.
formulation of relativistic mass.
While some people cathegorize the rst four of the above eects into relativistic kinematics,
the last one is termed as relativistic dynamics. We will discuss relativity from kinematical
aspects rst, then dynamical ones. In doing so, we need to reconsider Galilean transformation;
looking at what is wrong with it in the context of the universality of light and the unication
of space and time, and nding a way to x it using Lorentz transformation.
1.3 Galilean System versus Lorentz Transformation
It is natural to ask here that whether, assuming that a frame S

moving along the x-axis at


a speed of v relative to a frame S at rest, Galilean transformation equations below
y

= y z

= z x

= x vt t

= t (1.1)
originally applied to moving bodies in a mechanical framework of Newtonian physics, also hold
for electromagnetic radiation, that is, the propagation of electromagnetic waves in vacuum
derived from a complete set of Maxwells equations. As a consequence of the negative results
of the Michelson-Morley experiment, one must conclude that light travels in space at the same
speed in all direction, independent of the relative motion between the source, medium and
observer. This is equally qualitative with the invariance of the speed of light. In line with
this, time measurement is now in question; whether or not it is absolute.
In order to handle the diculties above, one must reconsider what is called as space;
whether it has the spatial components (x, y, z) only, or a temporal part t could be introduced
as the fourth component. Furthermore, whether space and time are actually interconnected
quantities directly obtained from space and time measurements. Thus, it is sensible to imagine
a world line as a line connecting one event to another in space-time (x, y, z, t) coordinates,
4 1. THE SPECIAL THEORY OF RELATIVITY
prompting that space and time are inseparable quantities. We mean an event as something
happening in a reference frame S in a specic location (x, y, z) and at a particular time t and
another event may exist in a reference frame S

at a point of (x

, y

, z

, t

). These two events


can be directly linked by a set of equations for the two frames of reference S and S

. In the
context of this, the Lorentz transformation equations below
y

= y z

= z x

= (x vt) t

= (t vx/c
2
) (1.2)
come into play. It is necessary to mention here that a factor is introduced to bridge classical
physics and relativistic approach. This factor is such that for moving objects at low speeds
(i.e. the speed of such objects is much smaller than the speed of light) reduces to one,
and hence relativity theory becomes classical mechanics. This factor is also something like a
self-contained quantity in that it provides an upper limit for the speed at which moving bodies
are able to travel. Mathematically, can be derived from various ways with the same result,
=
1
_
1 v
2
/c
2
(1.3)
With the help of both (1.2) and (1.3), more or less we now have an adequate tool needed to
relate one event in one inertial frame to another event in another inertial frame. However,
we intend to go through another route, which is more interesting and challenging. It is time
now to learn something new and dierent, but useful in understanding special and general
relativity theories.
1.4 The Four-Vectors
Although it is still possible to derive all formulations in the special theory of relativity without
the use of 4-vectors, we intend to introduce the concept of such vectors as they are extremely
helpful, and beautiful to some extent, in making descriptions of the theory much simpler and
far more transparent. One reason why we just introduce these to the second year students is
that relativity is better given for the rst time with no advanced techniques required. This
argument comes from historical perspectives in which special relativity is actually based on
purely physical, brilliant ideas that lead space and time to be a unied entity.
1.4.1 The position four-vector
We begin with contra-variant vectors for position dened as
x

(x
0
, x
1
, x
2
, x
3
) = (ct, x, y, z) (1.4)
1.4. The Four-Vectors 5
and co-variant vectors for position dened as
x

(x
0
, x
1
, x
2
, x
3
) = (ct, x, y, z) (1.5)
for which further formulations associated with them can be derived as follows.
dx

(dx
0
, dx
1
, dx
2
, dx
3
) = (c dt, dx, dy, dz) (1.6)
dx

(dx
0
, dx
1
, dx
2
, dx
3
) = (c dt, dx, dy, dz) (1.7)
Then, the inner product of (1.6) and (1.7) is written as
3

=0
dx

dx

dx

dx

= c
2
dt
2
(dx
2
+ dy
2
+ dz
2
) (1.8)
where we have here used Einstein summation convention. For cases with suciently nite
small changes, (1.8) can be written as
s
2
x

= c
2
t
2
(x
2
+ y
2
+ z
2
) (1.9)
where s
2
is called a space-time interval, or a geodetic line in space-time coordinates.
It is left for students to prove that this quantity is invariant under Lorentz transformation as
seen in (1.2). Here then, an interesting question is raised: what is the physical signicance of
s
2
to the dynamics of fast-moving bodies ? To this end, three cases below are considered.
Case 1: s
2
> 0, time-like region
Before further discussion, we again need to recall rst what is meant by an event. For now
it is sucient to say that an event is something happening in a specic location (x, y, z) and
at a particular time t in four-dimensional, space-time coordinates. This interval is normally
represented in the Minkowski diagram, a diagram for seeing how space-time coordinates
transform between dierent frames of reference (see Figure 1.1). This diagram consists of
two components: spacial position x

serves as a horixontal axis and temporal component ct,


or precisely ict, as a vertical axis. One of the relativistic consequences tells us that nothing
is faster than light. On the diagram, this is indicated by the fact that moving objects are
restricted only inside the light cone.
Let us now say for simplicity, the spacial part of space-time coordinates is only in the
x-direction, and hence (1.9) can be rewritten as
s
2
= c
2
t
2
x
2
> 0 (1.10)
6 1. THE SPECIAL THEORY OF RELATIVITY
x
ct
Figure 1.1: A light cone, showing past and future events in the Minkowski diagram
(taken from p.18, Ch.1, Relativity Demystied, David MacMahon, 2006).
The above relation represents regions in the diagram where v < c, making it possible to have
x

= 0 or x

= 0 calculated from Lorentz transformation for x

or x

. This means that


if two events are time-like separated, it is then possible to nd a frame of reference O

in which
the events occur at the same place.
Based on the invariance of s
2
under Lorentz transformation (1.2), measurements of time
interval in the frame of reference where the two events occur at the same place will result in
the proper time t
p
. This quantity is also dened as time interval measured by the observer
which is at rest relative to the events.
Case 2: s
2
< 0, space-like region
Again, the spacial part of space-time coordinates is assumed to be only in the x-direction,
and hence (1.9) can be rewritten as
s
2
= c
2
t
2
x
2
< 0 (1.11)
The above relation represents regions outside the time-like cone in the Minkowski diagram.
It is possible to have t

= 0 or t

= 0 calculated from Lorentz transformation for t

or t

.
It means that if two events are space-like separated, it is possible to nd a reference frame O

in which the events occur at the same time.


Based on the invariance of s
2
under Lorentz transformation, measurements of length, or
distance, in the frame of reference where the two events occur simultaneously will result in
the proper length, or distance, L
p
. This quantity is also dened as the length, or distance,
measured by the observer at rest relative to the events.
1.4. The Four-Vectors 7
Case 3: s
2
= 0, light-like region
This case is a special one in which the two events are said to be light-like separated. Then we
have
s
2
= c
2
t
2
x
2
= 0 (1.12)
and so c = x/t, implying that light propagates with a constant speed of c from one point
in the light line, or light path at the boundary of the time-like cone to another. For this case,
it is not possible to nd a frame of reference O

in which two events occur at the same place or


the same time. Thus, it is clear that both time dilation and length contraction as kinematical
aspects of relativity are derived from the rst two cases.
1.4.2 The energy-momentum four-vector
Special relativity theory may be classied into relativistic kinematics and relativistic dynamics.
The former deals with relativistic measurements of lengths, times, and speeds. It is basically
concerned with space-time coordinates only, and not with quantities like mass, force, energy,
and momentum. The latter, on the other hand, does deal with such quantities. Here, we will
derive the relativistic energy and momentum using the 4-vector formulation.
Let us begin with
p

(p
0
, p
1
, p
2
, p
3
) = (E/c, p
x
, p
y
, p
z
) (1.13)
for the energy-momentum contra-variant vectors, and its corresponding energy-momentum
co-variant vectors,
p

(p
0
, p
1
, p
2
, p
3
) = (E/c, p
x
, p
y
, p
z
) (1.14)
Then, the inner product of (1.13) and (1.14) will lead to
3

0
p

= p
0
p
0
+ p
1
p
1
+ p
2
p
2
+ p
3
p
3
=
E
2
c
2
(p
2
x
+ p
2
y
+ p
2
z
)
=
E
2
c
2
p
2
(1.15)
Note that, as is the case in (1.9), the quantity on the LHS of (1.15) must be invariant under
Lorentz transformation for energy-momentum. It is important at this stage to dene m
o
as
a rest mass of a moving object measured by an observer that is at rest with respect to the
object. This quantity is invariant under Lorentz transformation, meaning that its value is the
same for all rest observers in inertial frames. The invariance of m
o
, along with the universality
of the speed of light c, makes the LHS of (1.15) equal to m
o
c squared. Thus, (1.15) can be
8 1. THE SPECIAL THEORY OF RELATIVITY
written as
m
2
o
c
2
=
E
2
c
2
p
2
(1.16)
By rearranging the terms of the above equation, (1.16) can be rewritten as
E
2
= m
2
o
c
4
+ p
2
c
2
(1.17)
where E represents the total relativistic energy. A deeper look on (1.17) leads to two
possible values of energy,
E =
_
m
2
o
c
4
+ p
2
c
2
(1.18)
which means that the total energy of a particle with rest mass m
o
moving at a high speed v
may be positive or negative as a direc consequence of relativity. It was then Dirac in 1931 who
proposed a brilliant, although a bit speculative, idea that leads to a physical meaning of such
an equation. The two values of the total energy give rise to the existence of an anti-particle
with the same properties as its corresponding particle, except its electric charge. A detailed
explanation of anti-particles is given in the course of Introductory Nuclear Physics.
1.4.3 The rest energy and total relativistic energy
We have so far examined the use of energy-momentum four-vector to describe the dynamics
of moving bodies at high speeds with non-zero momentum. It is interesting to note that based
on Newtonian mechanics stationary objects bring no energy with them. We would like here to
examine the relativistic energy-momentum formulation for stationary objects by looking over
(1.13) and (1.14) when applied to the case of a non-moving object (equivalent to an object
having zero momentum) with rest mass m
o
as follows,
p

(p
0
, 0, 0, 0) = (E
o
/c, 0, 0, 0) (1.19)
p

(p
0
, 0, 0, 0) = (E
o
/c, 0, 0, 0) (1.20)
Then, we may replace (1.15) and (1.16) with
m
2
o
c
2
=
E
2
o
c
2
(1.21)
for which we can write
E
o
= m
o
c
2
(1.22)
as the rest energy of a particle with rest mass m
o
. With this, (1.17) can be rewritten as
E
2
= E
2
o
+ p
2
c
2
(1.23)
1.4. The Four-Vectors 9
For massless particles, (1.23) reads E = pc, a well-known energy-momentum relation for
photons. The detailed properties of photons will be formally given in a series of lectures on
the second part of this course. It is now time to nd what the relationship of E and E
o
is.
To do so we need to dene displacement four-vector as follows,
x

(x
0
, x
1
, x
2
, x
3
) = (c t, x, y, z) (1.24)
A simple but beautiful combination of (1.24) and t
p
= t/ yields velocity four-vector,
dened as
x

t
p
v

= (c, v
x
, v
y
, v
z
) (1.25)
multiplying both sides of (1.25) with m
o
results in
m
o
v

= ( m
o
c, m
o
v
x
, m
o
v
y
, m
o
v
z
)
p

= (E/c, p
x
, p
y
, p
z
)
(1.26)
which is exactly the same as relativistic momentum, previously given in (1.13). What is
beautiful here is that we have the zeroth component of (1.13) in the form of
E
c
= m
o
c (1.27)
from which we can write
E = mc
2
(1.28)
as the total relativistic energy. Here, we have used
m = m
o
=
m
o
_
1 v
2
/c
2
(1.29)
where m is dened as the relativistic mass of a moving particle (with rest mass m
o
) at a
high speed v. Please note that (1.29) tells us that it is impossible to accelerate an object of
rest mass m
o
to the speed of light, even with a continuous force acting on it. In other words,
the upper limit c of the speed of moving bodies never reaches.
1.4.4 The relativistic kinetic energy
It is of fundamental interest to examine the dierence between the total relativistic energy E
and the corresponding rest energy E
o
. We can write
E E
o
= mc
2
m
o
c
2
=
m
o
c
2
_
1 v
2
/c
2
m
o
c
2
= m
o
c
2
_
1
_
1 v
2
/c
2
1
_
(1.30)
10 1. THE SPECIAL THEORY OF RELATIVITY
In the non-relativistic regime where v is much less than c, the quantity in the bracket can be
expanded in terms of binomial series to the rst order, giving
E E
o
= m
o
c
2
_
1 +
v
2
2c
2
1
_
=
1
2
m
o
v
2
= K (1.31)
Surprisingly, the RHS of (1.31) is exactly the same as the non-relativistic kinetic energy
for a moving object of rest mass m
o
at constant speed v. Thus, we can write
E = E
o
+ K (1.32)
The above equation tells us that even if a physical body of rest mass m
o
is motionless with
K = 0 it still has an energy, that is, its rest energy E
o
. In this case, (1.32) becomes E = E
o
.
1.4.5 The relativistic Doppler eects
As with sound where there is a shift in frequency arising from the relative motion of the source
and observer, the frequency of light, or pulse, observed by an observer can be dierent from
that emitted by the source. This dierence is due to the relative motion of two frames; one
frame is a frame where the rest observer relative to the ground is standing, the other frame is
attached to the source. To analyze this phenomenon, we develop the following steps.
The power of Lorentz transformation previously written in (1.2) can be demonstrated here
by extending it to form energy-momentum transformation equations as follows,
E

= (E vp
x
) p

x
= (p
x
vE/c
2
) (1.33)
where we have assumed that an inertial frame S

is moving (along x or x

axis) at a speed of v
relative to an inertial frame S. Considering that m
o
c is invariant under Lorentz transformation,
(1.15) can thus be written, in the context of energy-momentum transformation, as
m
2
o
c
2
=
E
2
c
2
p
2
x
=
E
2
c
2
p
2
x
(1.34)
It is left for students to prove that the RHS of (1.34) is the same.
The beauty of Lorentz transformation in (1.2) can also be seen if we further explore (1.33)
by substituting these quantities,
E = E

p
x
= k
x
p

x
= k

x
1.4. The Four-Vectors 11
Figure 1.2: A longitudinal Doppler eect with the source moving towards an observer
(taken from Ch.10, Introductory Classical Mechanics, David Morin, 2003).
into (1.33). We then have

= ( vk
x
) k

x
= (k
x
v /c
2
) (1.35)
From the rst part of (1.35) we can derive the so-called relativistic Doppler shift for
frequency. To make it clear, let us consider a source of light in a reference frame S

emitting
ashes at a frequency
o
while moving at speed v directly toward us standing at the centre of
a reference frame S at rest. A simple question is that with what frequency do the ashes hit
our eyes ? Is it the same frequency as the original one, greater or smaller than the original
one ? What about if the source is directly moving away from us ? What about if the source is
moving across our eld of vision ? All of these questions are related to two eects of relativistic
Doppler eects, namely the longitudinal Doppler eect and the transverse Doppler eect.
In the following section, we will discuss the two eects separately.
Longitudinal Doppler eect
We solve the longitudinal Doppler eect as depicted in Figure 1.2 by writing k
x
= /c and
substituting it into the rst part of (1.35), giving

=
(1 v/c)
_
1 v
2
/c
2
=

1 v/c
1 + v/c
(1.36)
where

is the frequency of light-pulse emitted by the source and is the observed frequency.
Here,

serves as the proper frequency


source
. Thus for a source of light-pulse moving towards
an observer with speed v, the frequency measured by the observer in the ground is given by

receiver
=
source

1 + v/c
1 v/c
(1.37)
which is sensible in that the observed frequency
receiver
is greater than
source
.
12 1. THE SPECIAL THEORY OF RELATIVITY
Figure 1.3: A transverse Doppler eect with the source moving across an observer
(taken from Ch.10, Introductory Classical Mechanics, David Morin, 2003).
When the source moves away from the observer while still emitting the light-pulse, then
(1.36) becomes

=
(1 + v/c)
_
1 v
2
/c
2
=

1 + v/c
1 v/c
(1.38)
such that the longitudinal Doppler eect for this case is given by

receiver
=
source

1 v/c
1 + v/c
(1.39)
This is also sensible in that the observed frequency is lower than that of the source.
Transverse Doppler eect
Two cases of the transverse Doppler eect are shown in Figure 1.3, where the direction of
the source motion is taken to be parallel to the x-axis. In the rst case, we are dealing with
photons reaching our eyes when the source crosses the y-axis. In the second case, we see the
photons coming into our eyes along the y-axis. When we observe these photons the source
has been already at some distance from the y-axis.
As the propagation of light is perpendicular to the source motion, the observer in the rst
case will see the pulse at a higher frequency than the frequency emitted by the source. Thus,
in this case we can write

receiver
=

source
_
1 v
2
/c
2
(1.40)
In the second case, however, the source moves away from the y-axis while still emitting the
pulse along the y-axis. As a result, the observer will see the pulse at a lower frequency. Thus,
in the second case we can write

receiver
=
source
_
1 v
2
/c
2
(1.41)
1.4. The Four-Vectors 13
1.4.6 The wave operator
The relativistic position four-vector previously written in (1.4) and (1.5) can also be further
extended to create the four-dimensional wave operator as follows. Let us begin with

=
_

x
0
;

x
1
;

x
2
;

x
3
_
=
_

c t
;

x
;

y
;

z
_

=
_

x
0
;

x
1
;

x
2
;

x
3
_
=
_

c t
;

x
;

y
;

z
_ (1.42)
The inner product of the above equation gives

=
1
c
2

2
t
2

_

2
x
2
+

2
y
2
+

2
z
2
_
=
1
c
2

2
t
2

2
(1.43)
where
2
is well-known as the Laplacian operator. The four-dimensional wave operator
in (1.43) can be shortly written as
1
c
2

2
t
2

2

2
(1.44)
where
2
is the so-called dAlembertian operator. It is then left for students to prove that
this operator is invariant under Lorentz transformation in (1.2).
When the dAlembertian operator in (1.44) is applied to electric E and magnetic B elds,
then we have a set of wave equations representing the propagation of electromagnetic wave in
space-time coordinates as follows,

2
E =
1
c
2

2
E
t
2

2
E = 0
2
B =
1
c
2

2
B
t
2

2
B = 0 (1.45)
The above equation can be used to derive the invariance, or universality, of the speed of light
in vacuum. It follows that the speed of light in vacuum is independent of the relative motion
of the source and observer. For now, we do not understand yet what c in (1.45) looks like.
In order to solve what c is, we introduce here a famous set of Maxwell equations proposed
by Maxwell in 1865 that serve as a basis for the derivation of electromagnetic wave in vacuum.
The Maxwell equations are as follows,
. E =

o
. B = 0
E =
B
t
B =
o
E +
o

o
E
t
(1.46)
where denotes electric charge density per unit volume, and
o
, and
o
represent electric
permittivity, conductivity, and magnetic permeability in vacuum, respectively. Let us now
14 1. THE SPECIAL THEORY OF RELATIVITY
take a quick detour to derive electromagnetic wave equation in vacuum,
(B) =
o
(E) +
o

t
(E)
=
o

B
t

o

2
B
t
2
Applying the properties of vector identity to the LHS of the above equation, the equation
then changes into
(. B)
2
B =
o

B
t

o

2
B
t
2
Substituting one of (1.46) into the LHS of the above equation and assuming that vacuum has
zero conductivity, we then have

2
B
t
2

2
B = 0 (1.47)
representing space-time variations of magnetic eld B in vacuum. The same result as in (1.47)
for electric eld E is also obtained by taking the curl of E in (1.46) and then combining
with other ingredients to get

2
E
t
2

2
E = 0 (1.48)
Thus, it is proved that the electric and magnetic elds satisfy the wave equations.
Now, take a closer look on both (1.47) and (1.48) and compare them with (1.45). It is
very clear that the speed of light c in (1.45) is given by
c =
1

o
= 3 10
8
m/s
where
o
= 8.8510
12
C
2
/Nm
2
and
o
= 410
7
Ns
2
/C
2
. The result for the speed of light
in vacuum provides insight into the unique property of electromagnetic waves in the sense
that the waves, or light, always travel at the same speed c in vacuum. This value is the same
for all observers, regardless of their state of motion, or the relative motion of the source and
observer, leading to what we call the invariance of c in space-time coordinates. It is therefore
understood why in many physics books the speed of light is said to be universal.
However, it took many years to come for this insight to sink in before Einstein in 1905
used it as the second postulate of special relativity theory. This theory describes well many
diculties arising from the study of electricity and magnetism, and optics, experimentally and
theoretically established in the mid-19th century. It challenges our notions of what actually
meant by reality is. In his best times, Einstein did something like a tremendous leap forward,
leaving everything behind. With further advances in technology in the years since its discovery,
experimental evidence of the validity of the special relativity theory has really mounted.
1.5. Exercises 15
1.5 Exercises
All the following questions are taken from Modern Physics, 3rd edition, Serway et al., 2005.
1. A meter stick moving in a direction parallel to its length appears to be only 75 cm long
to an observer on the ground. Determine the speed of the stick relative to the observer.
2. A spacecraft moves at a speed of 0.9c relative to the earth surface. If its length is L
measured by a co-pilot on the spacecraft, what is its length measured by an observer at
rest with respect to the earth surface ?
3. A clock on a moving spacecraft runs a second slower per day relative to an identical
clock on earth. What is the relative speed of the spacecraft with respect to the earth ?
4. An astronaut at rest on earth has a heartbeat rate of 70 beats/min. What will be his
heartbeat rate when he is traveling in a spaceship at a speed of 0.9c as measured by
(a) an observer that is also on the spaceship ?
(b) an observer staying on earth ?
5. The average lifetime of a meson in its own reference frame is 0.026 s. If such a meson
moves at a speed of 0.95c, determine
(a) its mean lifetime measured by an observer on earth;
(b) the average distance it travels before decaying, as measured by an observer on earth.
6. Suzanne observes two light-pulses to be emitted from the same location but separated
in time by 3 s, while Mark sees the same pulses separated in time by 9 s.
(a) How fast does Mark move with respect to Suzanne ?
(b) According to Mark, what is the spatial separation of the two pulses ?
7. An electron has a kinetic energy 5 times greater than its rest energy. Determine
(a) its total energy;
(b) its speed.
8. Electrons in projection television sets are accelerated through a potential dierence of
50 kV.
(a) Calculate the speed of the electrons using the relativistic form of kinetic energy
assuming that the electrons start moving from rest;
(b) Calculate the speed of the electrons using the classical form of kinetic energy;
(c) Is the dierence in the speed between the two calculations signicant ?
16 1. THE SPECIAL THEORY OF RELATIVITY
9. Calculate, for the judge, how fast you were going in km per hour when you ran the red
light of wavelength 650 nm because it appeared Doppler-shifted green with a wavelength
of 550 nm to you.
10. A professor in physics staying on earth gives a nal exam to his students who are on a
spaceship traveling at a speed v relative to the earth surface. The moment the spaceship
passes the professor, he signals the start of the exam. If he wishes his students to have
a time t
o
(spaceship time) to complete all problems in the exam, show that he should
wait a time t (earth time) given by
t = t
o

1 v/c
1 + v/c
before sending a light-signal telling the students to stop doing the exam.
Chapter 2
THE GENERAL THEORY OF
RELATIVITY
The general theory of relativity is all about accelerated frames, or more precisely gravity.
In comparison to the special theory of relativity, the general relativity is mathematically more
complex; that is why it took many years for Einstein to come up with it after his astonishing
publications in 1905. The general relativity is a more fundamental theory of space and time;
it describes gravity as the distortion of the structure of space and time caused by the presence
of massive objects. Here, instead of providing a detailed mathematical framework of general
theory, this chapter will only give a avour for the subject in question in a fashionable way,
in the context of getting a sense of the nature of general relativity without having master to
the theory. This chapter will therefore not include lengthly derivations or detailed physical
explanations; the goal is just to make general relativity easier to learn and understand, and
more accessible to physics students.
2.1 The Equivalence Principle
If we consider two cases, in relation to force and weight, where mass of an object is involved,
we realize that mass seems to have two dierent properties, that is, an inertial mass describing
a reluctance to accelerate and a gravitational mass showing a response to a gravitational eld,
or gravity. There is no fundamental reason actually to justify that the dynamics governing
the resistance of an object to move (when an external force is applied) is likely equivalent to
the property governing how hard gravity pulls on that object. It may be useful in the rst
place to write m
i
for denoting an inertial property and m
g
for a gravitational property. It is
then interesting to evaluate these two concepts concerning mass of a body; whether they are
the same, and to examine its consequences upon the dynamics of a moving body. We are now
in a position ready to go for further studying this question.
17
18 2. THE GENERAL THEORY OF RELATIVITY
Figure 2.1: The principle of equivalence between two frames of reference where Bakwan
and Afdol are located (taken from p.44, Ch.2, Modern Physics, Randy Harris, 2007).
From the above points of view, general relativity given in this course simply begins with
the so-called Einsteins Equivalence Principle as follows :
Inertial mass m
i
, which appears in the Newtons second law of motion in the form of
F = m
i
a, is equal to gravitational mass m
g
, which appears in the formulation of bodys
weight in the form of W = m
g
g. Thus for F = W, a body of mass m under the inuence
of an external force is accelerated by a = (m
g
/m
i
) g. The ratio m
g
/m
i
is then under
investigation. Experimental evidence from various materials used conrms that there is
no signicantly numerical dierence in the ratio.
Let us consider two situations in which Bakwan and Afdol are located (see Figure 2.1).
While Bakwan stays at rest in a small box under the inuence of a gravitational eld,
Afdol is accelerating upward with uniform acceleration, say g, in a gravity-free space.
Here, the equivalence principle says that there is no local experiment they can perform
that will tell them which of the two situations they are in. Thus, the two situations are
completely equivalent.
Based on the equivalence principle above, Einstein in 1916 proposed another two famous
postulates, namely the postulates of general relativity as follows :
The laws of physics are in the same form and hold for all observers, independent of their
state of motion; whether they are in inertial frames or non-inertial frames.
It is not possible to distinguish between gravity and acceleration, in the sense that in
the absence of gravitational eects, a physical event occurring in an accelerated frame
could also occur in the same manner just as it is in an inertial frame under the inuence
of a gravitational eld.
2.2. Consequences of the Equivalence Principle 19
Figure 2.2: Two events, namely dropped objects are physically equivalent in the frames
where Bakwan and Afdol are standing (taken from p.44, Ch.2, Modern Physics, Randy
Harris, 2007).
The above postulates are well described in Figure 2.2. In the linearly accelerating frame
where Afdol is standing, a dropped object appears to accelerate downward just as it is in the
inertial frame where Bakwan is standing. Thus, it does not really matter where ever you are;
whether you are in an inertial frame or a non-inertial frame, the same law of physics applies.
2.2 Consequences of the Equivalence Principle
The principle of equivalence, along with the postulates of the general theory of relativity,
has striking consequences regarding with the nature of space and time in the presence of
gravity. Three astonishing predictions derived from the principle of equivalence have attracted
particular attention to scientic communities at the time general relativity was launched
in 1916. The predictions include: (1) time dilation due to gravity; (2) gravitational red-shift;
and (3) the deection of light by the sun. From these predictions, we can see the elegance of
the general relativity. As this theory requires complex mathematics with tensor calculus and
dierential geometry involved, we here only provide qualitative discussion for each prediction.
2.2.1 Gravitational time dilation
An interesting eect predicted by the general relativity theory is that time is altered by gravity.
To discuss this, let us consider the following two scenarios :
A source of light positioned at the top of a building tower produces ashes at time
intervals t
source
. An observer on the ground receives the ashes at time interval t
receiver
(see sketch on the left of Figure 2.3).
20 2. THE GENERAL THEORY OF RELATIVITY
Figure 2.3: The propagation of light pulses are physically equivalent in the left frame
where gravity exists and in the right frame where acceleration (with no gravity) works
(taken from Ch.13, Introductory Classical Mechanics, David Morin, 2003).
A rocket with its bodys length h moves upward with constant acceleration g in a
gravity-free space. A source of light on the top end of the rockets body emits ashes at
time intervals t
source
. An observer at the bottom end of the rockets body receives the
ashes at time interval t
receiver
(see sketch on the right of Figure 2.3).
What do you think about t
source
and t
receiver
in both cases ? If we look at back previous
discussions, then we have to say that the two scenarios have to be the same. We can therefore
examine the second scene to solve what is happening in the rst scene.
Now, let us stay focus on the second scene by considering a series of quick pulses emitted
by the light source on the top of the rockets body with very small time interval t
source
such
that many pulses are released before the rocket moves considerably. By the time the pulses
reach the receiver, the rocket and hence the receiver will be moving. Assuming that the speed
of the rocket v is very small compared to the speed of light c, we have from the classical
Doppler eect the relationship between the frequency produced by the source
source
and the
observed frequency
receiver
as follows,

receiver
=
_
c + v
c
_

source
=
_
1 +
v
c
_

source
(2.1)
Here at this stage, the story begins. What we really want is to include the distance h between
the source and the observer in the Doppler eect given by (2.1).
Notice that the observer at the bottom receives the light after a time t = h/c. Accordingly,
by this time she or he is moving upward at speed v = gh/c (assuming that the rocket moves
from rest with zero initial velocity). Thus, the light is observed in a frame of reference that
moves at speed v = gh/c towards the frame of reference (i.e., the source) in which it was
emitted. From this point of view, it is understood that the dynamics of the classical Doppler
eect is merely determined by the height at which the source of light is located (measured
2.2. Consequences of the Equivalence Principle 21
from the position where the receiver stays). We can therefore write

receiver
=
_
1 +
gh
c
2
_

source
(2.2)
and the corresponding equation for gravitational time dilation is then given by
t
source
=
_
1 +
gh
c
2
_
t
receiver
(2.3)
where we have used 1/
source
= t
source
and 1/
receiver
= t
receiver
directly from (2.2) to
get (2.3). The above equation tells us that, in the context of the rst scene in Figure 2.3, an
observer on the ground will see the clock at the top of the tower running faster by a factor
1 + gh/c
2
than its identical clock held by the observer. This is a truly time dilation eect,
but it has nothing to do with the relative motion between the source and the observer, as
previously discussed in Chapter 1. This time dilation eect gives a very small dierence in
the time measurement performed at the Earth surface and at some altitude above the ground.
However, this time dierence is made possible with the help of high-precision clocks based on
atomic oscillations of extremely short period.
2.2.2 Gravitational red-shift
An interesting phenomenon could be simply analyzed when positions of the light source and
the observer in both scenes of Figure 2.3 are exchanged. In this new scenario, the light pulses
are emitted from the source located at the bottom end of the rockets body (or on the ground)
while the observer is positioned at the top end of the rockets body (or on the top of the tower).
In this case, the gravitational time dilation becomes
t
source
=
_
1
gh
c
2
_
t
receiver
(2.4)
and the corresponding shift in frequency is given by

receiver
=
_
1
gh
c
2
_

source
(2.5)
The above equation tells us that, for a beam of light rays, the frequency of the light decreases,
or its wavelength grows longer, as the beam rises from the ground to some height above under
the inuence of a gravitational eld. This phenomenon is known as gravitational red-shift,
implying that the frequency of atomic radiation in the presence of a strong gravitational eld
is shifted to a lower frequency when it is compared with the same emission in the presence of
a weak gravitational eld.
22 2. THE GENERAL THEORY OF RELATIVITY
Figure 2.4: The path of a light beam sending horizontally in an accelerating frame
with no gravity is deected downward seen by the observer in the frame. This bending
of light is exactly the same as it is in an inertial frame where gravity exists, as shown in
the right picture (taken from p.50, Ch.2, Modern Physics, Randy Harris, 2007).
2.2.3 The bending of light rays by gravity
The fact that gravity is responsible for a red-shift phenomenon is a clue for an amazing story,
that is, gravity is able to aect light. But, how could that be ? Doesnt light have zero mass,
and by this it should not be never aected by gravity because the way in which gravity works
on an object is to pull its mass m. Likewise, the Newtons universal law of gravitation requires
the presence of a mass of a body in a gravitational eld. Hence, in the absence of mass (for
the case of light, or photons in general) the classical law of gravitation is no longer relevant.
Einstein answered this puzzle by suggesting that the Newtons formulation for gravity is only
correct for a weak eld and holds only for a Euclidean space, in which light will always take
the shortest path when traveling from one point to another. It follows that in this geometry
light travels in a straight line, and this line lies on a two-dimensional plane.
However, in the presence of a strong eld the path taken by light rays may not be straight.
The geometry used to describe the light path near a massive object is a non-Euclidean space
developed from the general theory of relativity. Based on this new paradigm, the geometry of
the four-dimensional space-time under strong gravity is curved. Thus, light is deected when
it passes a region where the regularity of space-time is disturbed by a strong gravitational
eld. It can be seen from the third picture of Figure 2.4 that a ray of light in an accelerating
frame where gravity is negligible would appear to bend downward when viewed by the observer
within the frame. Based on the equivalence principle between the third and fourth pictures
of Figure 2.4, the light is then predicted to bend downward by a gravitational eld.
2.3. Exercises 23
Figure 2.5: The bending of a light ray from a star, showing gravitational deection due
to the presence of the Sun as a massive object (taken from p.56, Ch.2, Modern Physics,
Serway et al., 2005).
Evidence for this theoretical prediction has been found in an experiment using a beam of
laser projected horizontally to a distance of 6000 km, where a deection of less than 1 cm was
observed for such a distance. Although this path deection is suciently small compared to
the total distance the light has traveled, this bending shows the nature of the passage of space
and time in the presence of a strong gravitational eld. A natural, striking example for this
phenomenon comes from a light ray of a star when passing near the Sun (see Figure 2.5). The
starlight was deected in the curved space-time triggered by the Suns massive mass during
a total solar eclipse in 1919. When this discovery was announced to scientic communities
the prediction based on the general theory of relativity was conrmed, and hence Einstein
became an outstanding scientist.
2.3 Exercises
1. With regard to reference frames, how does general relativity dier from special relativity?
(taken from Ch.2, Modern Physics, 3rd edition, Serway et al., 2005).
2. Two identical clocks are in the same house, one is placed upstairs in a bedroom and the
other is in the kitchen. Which clock runs more slowly ?
(taken from Ch.2, Modern Physics, 3rd edition, Serway et al., 2005).
3. A plane ies at constant height h. What should its speed be so that an observer on the
ground sees the planes clock tick at the same rate as a ground clock ? Assume that the
speed v of the plane is much small compared to the speed of light c.
(taken from Ch.13, Introductory Classical Mechanics, David Morin, 2003).
4. A clock starts on the ground and then moves up a tower at constant speed v. It sits on
the top for a time T and then descends at constant speed v. If the tower has height h,
24 2. THE GENERAL THEORY OF RELATIVITY
how long should the clock sit at the top so that it comes back showing the same time
as a clock that remained on the ground ? Assume that the speed v of the plane is much
small compared to the speed of light c.
(taken from Ch.13, Introductory Classical Mechanics, David Morin, 2003).
5. A person B moves at speed v which is much small compared to the speed of light c in a
circle of radius r around another person A. By what fraction does Bs clock run slower
than As ? Calculate this in three ways as follows :
(a) As frame;
(b) the frame whose origin is B and whose axes remain parallel to an inertial set of axes;
(c) the rotating frame centred at A and rotates around A with the same frequency as
that of B.
(taken from Ch.13, Introductory Classical Mechanics, David Morin, 2003).
Chapter 3
THE BLACK-BODY RADIATION
The achievement made in theoretical and experimental physics towards the end of the-19th
century had been immense. Classical mechanics and thermodynamics were well understood.
The concept of electromagnetics was also remarkably established. It was found that light is a
form of electromagnetic waves, providing a rm theoretical framework for the wave theory of
light, which could account for most phenomena in optics. Thus, most physicists at that time
believed that the combination of the three subjects could account for all physical phenomena.
In fact, it was a period of a turmoil state when there were surprising experimental results,
which based on classical theory were totally inexplicable. One dilemma lay in the observations
of thermal radiation. Existing classical theory was unable to explain the observed frequency
(or wavelength) of radiant energy. The incredibly greatest minds of physics in the period
were about to begin. Out of the turmoil state came a new philosophy of science a new
way of thinking the so-called Quantum Physics. The way we think about natural laws
particularly in microscopic world has totally changed since then. In this new paradigm, light
is considered not only as a wave but also as a series of bundles of energy called quanta.
Unlike Newtonian mechanics where dynamic variables, such as position and momentum can
be accurately determined with high precision, quantum world makes these variables uncertain
in that all measurements made are undeterministic.
This chapter is aimed to examine the birth of quantum physics that has shaped the world
of physics. We rst begin with experimental observations of the spectrum of thermal radiation,
what comes later to be known as black-body radiation, that put physicists at that time
into a Pandora Box. The second issue to discuss is the theoretical work of Wien in 1893, who
provided a primitive formula for the distribution of radiant energy. In line with this, we then
discuss the contributions of Rayleigh-Jeans in 1900 to the problem in question. Finally, we
revisit the notion of the photon concept of electromagnetic radiation in the light of Plancks
ideas proposed in 1901. This great work of Planck, along with relativity theory suggested by
Einstein, serves as a connecting bridge between classical physics and modern physics.
25
26 3. THE BLACK-BODY RADIATION
i
n
t
e
n
s
i
t
y

i
n

w
a
t
t
/
m
2
(a) (b)
Figure 3.1: Plots of radiant energy E, showing the characteristics of thermal radiation
of a black-body as a function of wavelength : (a) at various temperatures T and (b) at a
given temperature with theoretical predictions suggested by Rayleigh-Jeans and Planck
(taken from Ch.1, Quantum Physics, Stephen Gasiorowicz, 1996).
3.1 Thermal Radiation
In 1859, Kircho proposed a theorem regarding with the relation between the coecients
of emission and absorption of electromagnetic radiation; how these could be related to the
spectrum of thermal equilibrium radiation, the so-called black-body radiation phenomenon.
He challenged the community to work this out. A black-body is an object that absorbs all
the radiant energy falling on it and hence the black-body reects no light, for which it would
appear as black. A black-body is thus a perfect absorber, as well as a perfect emitter. Kircho
suggested a functional dependence of the radiant energy E on temperature T and frequency
(or wavelength ),
E = E (T, ) = E (T, ) (3.1)
It has since then been of fundamental interest among physicists during the last two decades of
the-19th century to nd out what the explicit form of the function is. This has been examined
through an approach of both experimental and theoretical considerations.
In 1879, Stefan based on experimental results argued that the radiant energy emitted by a
hot body is proportional to the fourth power of its absolute temperature. The same conclusion
based on classical thermodynamics was derived by Boltzmann in 1884. The ndings led to
the Stefan-Boltzmanns law. Here, we are not going to provide the derivation of the law.
3.2. Wiens Displacement Law 27
Rather, we pick up the result as follows,
E T
4
(3.2)
where E denotes the energy emitted per unit time per unit area from the surface of a black-body
at a given temperature T, and is the Stefan-Boltzmann constant (5.67 10
8
Wm
2
T
4
).
3.2 Wiens Displacement Law
In 1893, Wien proposed a formula to answer to the question previously posed by Kircho.
Wiens displacement law was derived using a combination of classical electromagnetics
and thermodynamics, as well as dimensional analyses. From the two basic concepts, it can be
shown that
T
1
(3.3)
The above relation says that the wavelength of a set of radiant waves changes inversely with
absolute temperature T, as seen in Figure 3.1(a).
Wien then went further by combining his own law with the Stefan-Boltzmann law. If we
dene E() as the energy density radiated, then we can write
E() T
5

5
(3.4)
for which we can derive
E =
5
f(T) =
3
g(/T) (3.5)
where both f = f(T) and g = g(/T) are constants. The resulting prediction (3.5) matches
experimental observations, particularly for a region of small wavelengths (or high frequencies),
but it was brokendown for a region of large wavelengths (or low frequencies). Equation (3.5)
is the complete form of Wiens law and sets constrains on the black-body radiation spectrum.
But it is actually incomplete because f = f(T) and g = g(/T) are both undetermined.
The complete form of the energy density of spectral distribution cannot be determined from
classical physics.
3.3 Rayleigh-Jeans Law
In 1900, Rayleigh analysed experimental data of black-body radiation and its corresponding
theory. His interest was stimulated by the inadequacies of Wiens law in (3.5) in accounting
for the large-wavelength (or low-frequency) behaviour of black-body radiation as a function
28 3. THE BLACK-BODY RADIATION
of temperature. The reason for the inclusion of Jeans name in what became known as the
Rayleigh-Jeans law is that there was an error in Rayleighs analysis, which was corrected
by Jeans in 1906. Based on theoretical considerations, Rayleigh-Jeans found that the spectral
energy density radiated is given by
E =
8
2
c
3
E (3.6)
where E is something that must be in a unit of energy. Here Rayleigh put E = kT based on
the principle of the equipartition of energy from classical kinetic theory into (3.6). Hence, we
have
E =
8
2
c
3
kT (3.7)
for the energy spectral distribution at a given temperature. This formula is only valid at low
frequencies (or large wavelengths). It does not hold for high frequencies (or small wavelengths)
because the spectrum of black-body radiation does not increase as
2
to innite frequency, as
shown in Figure 3.1(b).
Thus, both Wiens law and Rayleigh-Jeans law are counter-part in that they complete one
to another. However, what people want is a comprehensive theory that can be used to explain
the black-body radiation for all ranges of frequency or wavelength. It seems that the existing
theory based on classical view of thermal radiation of a black-body where such radiation is
considered to propagate in space as waves is no longer relevant.
3.4 Plancks Formula
Here we do not intend to derive explicitly what Planck did in his best times. Rather, we try to
connect (3.6) previously derived by Rayleigh to the correct formula. From classical statistics,
E is related to the mean energy of a harmonic oscillator having two degrees of freedom, which
should be
1
2
kT +
1
2
kT = kT. This is what exactly Rayleigh did in deriving his law, which
turns out to be the correct expression for the black-body radiation law at low frequencies (or
large wavelengths), as shown in Figure 3.1(b).
An interesting question is that why did Planck not make such substitution for E ? Firstly,
the equipartition theorem of Maxwell-Boltzmann is a result of statistical thermodynamics,
which was a point of view that he had rejected. Secondly, the Maxwells kinetic theory could
not account for the problem of specic heats of diatomic gases at high temperatures. Thirdly,
Planck did not agree fully with the Boltzmanns statistical approach.
In 1901, Planck worked out the problem of fundamental interest the energy spectral
distribution of the black-body radiation. Planck has introduced the concept of quantisation
by assuming that radiation consists of a bundle of quanta each having an energy which is
proportional to the radiant frequency . The energy density distribution E of the black-body
3.4. Plancks Formula 29
radiation can then be written as a function of radiant frequency as follows,
E() =
8 h
3
c
3
1
e
h/kT
1
(3.8)
The above expression satises all regions of frequency and wavelength in the spectrum, as
seen in Figure 3.1(b). The Plancks formula in (3.8) for the black-body radiation is valid
over the whole spectrum of wavelengths, or accordingly frequencies. It covers the empirically
observed ndings of the spectral distribution of electromagnetic radiation that are previously
unexplained by classical theories of both electrodynamics and thermodynamics.
It is straightforward to integrate (3.8) to nd the total energy density of radiation U in
the black-body spectrum,
U =
_

0
E() d =
8 h
c
3
_

0

3
e
h/kT
1
d (3.9)
for which the total power of radiation emitted per unit area can be calculated from classical
theory of radiation, I = c U/4, such that we can write
I =
2k
4
T
4
h
3
c
2
_

0
(h/kT)
3
e
h/kT
1
d
_
h
kT
_
=
2k
4
T
4
h
3
c
2
_

0
x
3
e
x
1
dx (3.10)
where I denotes the total energy radiated per unit time per unit area and x h/kT. The
integral can be evaluated with the help of some steps below (see p.5, Gasiorowicz, 1996),
_

0
x
3
e
x
1
dx =
_

0
x
3
dx e
x

n=0
e
nx
=

n=0
1
(n + 1)
4
_

0
y
3
e
y
dy = 6

1
1
n
4
=

4
15
Equation (3.10) then becomes
I =
_
2
5
k
4
15h
3
c
2
_
T
4
= T
4
(3.11)
where the quantity in the bracket is replaced with , and hence
=
2
5
k
4
15h
3
c
2
(3.12)
representing the Stefan-Boltmann constant, as previously shown in (3.2).
Most people at the time Planck proposed his ideas in verifying the experimental data of
black-body radiation got an impression that the Plancks concept of energy quantization was
only for the case of thermal radiation. It remained until Einstein who used the work of Planck
to show that light is quantized in the case of photo-electric eect.
30 3. THE BLACK-BODY RADIATION
3.5 Exercises
1. Show that the Rayleigh-Jeans spectral distribution of black-body radiation given in (3.7)
is of the form required by Wiens law given in (3.5).
(taken from Ch.2, Introductory Quantum Mechanics, Richard Libo, 1980).
2. Obtain the correct form Wiens undetermined function f(T) from Plancks formula
given in (3.8).
(taken from Ch.2, Introductory Quantum Mechanics, Richard Libo, 1980).
3. Use (3.8) to prove Wiens displacement law in the form of

max
T = constant
(taken from Ch.2, Introductory Quantum Mechanics, Richard Libo, 1980).
4. (a) Use the Stefans law to calculate the total power radiated per unit area by a tungsten
lament at a temperature of 3000 K (assume that the lament is an ideal radiator).
(b) If the tungsten lament of a lightbulb is rated at 75 W, what is the surface area of
the lament ? (assume that the main energy is lost due to radiation).
(taken from Ch.3, Modern Physics, 3rd edition, Serway et al., 2005).
5. At what wavelength does the human body emit the maximum electromagnetic radiation?
(assume that the skins temperature is about 70

F).
(taken from Ch.3, Modern Physics, 2nd edition, Randy Harris, 2007).
Chapter 4
THE PARTICLE PROPERTIES OF
ELECTROMAGNETIC RADIATION
As discussed in Chapter 3, Planck proposed a remarkable formula for the spectral distribution
of black-body radiation, which shed light on the photon concept of quantisation. There soon
followed an understanding of the quantum nature of electromagnetic radiation. However, it
was not until Einstein who suggested that the concept was applied to real cases when metals
were irradiated with light. Another support for the concept came from experimental evidence
of the particle nature of photons observed by Compton. These two fascinating phenomena
contribute signicantly to the development of modern physics. One of the corner stones of
quantum physics is wave-particle duality, meaning that things may behave as waves or
particles depending upon a physical situation. In a classical situation, say the propagation
of sunlight in space for example, such electromagnetic radiation is considered as waves with a
continuous ranges of energy spread out. In this chapter, however, we study the complementary
topic electromagnetic radiation behaves as a collection of discrete particles.
4.1 Photo-Electric Eect
An important contribution to the energy quantisation of electromagnetic radiation came from
the work of Einstein who in 1905 applied the concept of the quantum nature of light to explain
unresolved, experimental ndings in the photo-electric eect originally discovered by Hertz
in 1887. Hertz found peculiar properties of metals while observing his famous experiments on
electromagnetic waves to validate the Maxwells hypothesis of the speed of light.
The experimental set up of the photo-electric eect experiments is depicted in Figure 4.1,
where a metal plate of a certain material placed as a cathode in the apparatus is irradiated with
light of a given frequency . There were a number of surprising results from the experiments.
The most remarkable feature was that the kinetic energies of the photo-electrons emitted from
31
324. THE PARTICLE PROPERTIES OF ELECTROMAGNETIC RADIATION
Figure 4.1: Sketch of the experimental set up of photo-electric eect experiments (taken
from Ch.2, Introductory Quantum Mechanics, Richard Libo, 1980).
the metal surface are independent of the intensity of the incident radiation, but are linearly
dependent upon the incident frequency (or wavelength). This feature could not be explained
by classical theory of electromagnetic radiation at that time.
Einstein proposed a brilliant idea to solve the problem in question by assuming that the
radiation consists of a collecting bundle of energy called quanta of the same energy h. When
the bundle of energy is absorbed by the metal, electrons may receive sucient energy to escape
from the metal surface, against the energy that binds them. When the intensity of the incident
light is increased more electrons will be ejected from the surface, with their kinetic energies
remaining unaltered although the associated photo-electric current measured by Ammeter is
larger, as shown in Figure 4.2(a). The result for an increase in the photo-current as the light
intensity increases is actually predicted from classical view of electromagnetic radiation. But,
the result for the independence of kinetic energy on the light intensity is totally unexpected.
This is a point where the glorious story of the quantum nature of light begins.
To cope with the unexpected results above, Einstein further suggested that the maximum
kinetic energy K
max
of the photo-electrons linearly depends on both the incident frequency
(or wavelength) and the work function W of a given material, as shown in Figure 4.2(b).
The work function of a metal here is dened as the minimum amount of work necessary to
remove the electrons from the metal surface. Thus, this kinetic energy is expressed as
K
max
= h W
=
hc

W
(4.1)
where the term max is used to show that W is needed to free the least strongly bond electrons.
4.2. Compton Eect 33
Figure 4.2: Some of the surprising results for experiments of photo-electric eect,
showing (a) photo-current is only aected by the intensity of incident light with a xed
stopping potential for a given material; and (b) the dependence of the maximum kinetic
energy of photo-electrons on both the incoming frequency and the material used for a
cathode-plate (taken from Ch.3, Modern Physics, Serway et al., 2005).
Determination of the magnitude of the work function for a given material involves placing
a photo-cathode in an opposing potential so that, when the potential reaches a certain value,
the ejected electrons can no longer reach a collecting plate that serves as an anode, causing
photo-electric currents to fall to zero. This situation occurs at a value of an applied voltage
called the stopping potential V
s
. In this situation, K
max
= eV
s
such that (4.1) becomes
eV
s
= h(
o
) (4.2)
where we have used W = h
o
with
o
being the threshold frequency, that is, the minimum
frequency needed to eject the electrons from the metal surface. Note that based on (4.2) eV
s
must be a linear function of the incident frequency, with the slope of the straight line being h
known as the Plancks constant experimentally found to be 6.63 10
34
Js, independent of
the nature of the material chosen.
The central point in Einsteins contribution to the photo-electric eect experiments is that
electromagnetic radiation behaves as a collection of particles, each having a discretized energy
as opposed to the classical view of continuous wave energy for the radiation. With all respects,
Einsteins explanations agree well with experimental evidence, for which the work earned him
the Nobel Prize in physics in 1921.
4.2 Compton Eect
In 1922, Compton provided direct experimental evidence for the correctness of the photon
concept of electromagnetic radiation. In the experiment known as the Compton scattering,
344. THE PARTICLE PROPERTIES OF ELECTROMAGNETIC RADIATION
Figure 4.3: Sketch of the Compton eect, showing (a) initial and (b) nal states of
the X-ray scattering experiment (taken from Ch.2, Introductory Quantum Mechanics,
Richard Libo, 1980).
he used a beam of X-ray of initial energy h and momentum p to bombard a targeted electron
initially at rest, as shown in Figure 4.3(a). It was found that the X-ray beam was scattered in
a manner which is not consistent with classical theory of electromagnetic radiation. According
to the classical wave theory, the electron would scatter electromagnetic energy in all directions
at the same frequency as the incoming radiation frequency. But in fact, the scattered X-ray
beam has a dierent, outgoing frequency. To this end, Compton argued that the incoming
radiation should be treated as a beam of photons with individual photons scattering elastically
o individual electrons.
Based on the sketch of the Compton eect experiment (see Figure 4.3) where both energy
and momentum must be conserved, the conservation of energy is then given by
h + m
o
c
2
= h

+ m
o
c
2
(4.3)
where m
o
is the rest mass of electron, and

are the frequencies of incoming and scattered
X-ray beam, respectively, and is dened to be (1
2
)
1/2
, where = v/c. Note that
here we use the relativistic expression for energy, and later momentum, as the recoil electron
is frequently observed to move very fast after the collision. We do not need to worry about
the expression for photons as the non-relativistic expression for objects that always move at
speed c does not exist.
4.2. Compton Eect 35
Unlike energy, conservation of linear momentum is separately given for each direction, one
component in the direction of the incident radiation and the other component perpendicular
to the former. The equations denoting the momentum conservation are
h

= m
o
c cos +
h

cos (4.4)
and
0 = m
o
c sin +
h

sin (4.5)
where is an angle through which the photon is scattered and is an angle of the recoil
electron, kicked o from its original position as seen in Figure 4.3(b).
Compton pointed out that the scattered beam has a larger wavelength than the incident,
associated with less amount of energy for the scattered beam. The dierence in wavelength
between the two beams can therefore be directly derived from the energy and momentum
conservation, and is measured as an increase in wavelength of the incoming radiation in
a billiard-ball type of collision with a stationary electron. We can then write
=

=
h
m
o
c
(1 cos ) (4.6)
where and

denote the wavelengths of the incoming radiation and scattered photon,


respectively, and h/m
o
c = 0.024

A is called the Compton wavelength. Thus, the dierence
in the wavelength between the incident and scattered photons depends only on the angle
of the scatter.
Let us take a closer look at (4.6). If the targeted electron is replaced with an atom, for
example, then m
o
would be the mass of the atom, which is much more larger than the electron
mass. Consequently, the shift in the wavelength is suciently small to observe so that it
is eectively zero. This situation is similar to the case of an elastic collision between a small
projectile and a relatively massive body as a target. After a head-on collision, the projectile
is reected back from the body with the same values of both kinetic energy and momentum,
but is in the opposite direction. For this case, the increase in wavelength is maximum for
which this type of scattering is called backward scatter of the incoming photon with = 180

because the collision imparts the maximum possible energy to the recoil electron.
The measurements of the shift in the wavelength and the kinetic energy of the recoil
electron are in good agreement with the Comptons theoretical predictions. These provide
convincing justication of the correctness of the classical, billiard-ball collision interpretation,
demonstrating the particle behaviour of the photon. The results for the Compton scattering
experiments opened up the problems of classical view of electromagnetic radiation theory, for
which Compton won the Nobel Prize in 1927.
364. THE PARTICLE PROPERTIES OF ELECTROMAGNETIC RADIATION
4.3 Exercises
1. The photo-electric threshold of tungsten is 2300

A. Determine the kinetic energy of the
electrons ejected from the metal surface by ultraviolet light of wavelength 1900

A.
(taken from Ch.2, Introductory Quantum Mechanics, Richard Libo, 1980).
2. The work function of zink is known to be 3.6 eV. What is the energy of the most energetic
photo-electron emitted by ultraviolet light of wavelength 2500

A ?
(taken from Ch.2, Introductory Quantum Mechanics, Richard Libo, 1980).
3. Photo-electrons emitted from a cesium plate that are illuminated with ultraviolet light
of wavelength 2000

A are stopped by a potential of 4.21 V. What is the work function
of cesium ?
(taken from Ch.2, Introductory Quantum Mechanics, Richard Libo, 1980).
4. Ultraviolet light of wavelength 3500

A falls on a potassium plate. The maximum kinetic
energy of the photo-electrons is 1.6 eV. What is the work function of potassium ?
(taken from Ch.1, Quantum Physics, Stephen Gasiorowicz, 1996).
5. Consider the following metals lithium, beryllium, and mercury, having work functions
of 2.3 eV, 3.9 eV, and 4.5 eV, respectively. If light of wavelength 300 nm is incident on
each of these metals, then determine
(a) which metal exhibit the photo-electric eect;
(b) the maximum kinetic energy for the photo-electrons in each case
(taken from Ch.3, Modern Physics, Serway et al., 2005).
6. A photon of energy h collides with a stationary electron of rest mass m
o
. Show that it
is not possible for the photon to impart all its energy to the electron.
(taken from Ch.2, Introductory Quantum Mechanics, Richard Libo, 1980).
7. A 100 MeV photon collides with a proton at rest. What is the maximum possible energy
loss for the photon ?
(taken from Ch.1, Quantum Physics, Stephen Gasiorowicz, 1996).
8. A 100 keV photon collides with an electron at rest. It is scattered at 90

. What is its
energy after the collision ? What is the kinetic energy in eV of the electron after the
collision, what is the direction of the recoil electron ?
(taken from Ch.1, Quantum Physics, Stephen Gasiorowicz, 1996).
4.3. Exercises 37
9. A beam of X rays is scattered by electrons at rest. What is the energy of the Xray if
the wavelength of the X-ray scattered at 60

to the beam axis is 0.035



A ?
(taken from Ch.1, Quantum Physics, Stephen Gasiorowicz, 1996).
10. Gamma rays (high-energy photons) of energy 1.02 MeV are scattered from electrons
that are initially at rest. If the scattering is symmetric, that is, if = then determine
(a) the scattering angle ;
(b) the energy of the scattered photons
(taken from Ch.3, Modern Physics, Serway et al., 2005).
384. THE PARTICLE PROPERTIES OF ELECTROMAGNETIC RADIATION
Chapter 5
THE BOHR MODEL OF THE ATOM
5.1 The Early Atomic Model
The rst half of the 20th century was generally considered as a period of remarkable advances
in modern physics, pioneered by the triumphs of relativity and quantum theories, as well as
similar discoveries in the elds of atomic and nuclear physics. As with relativity theory and
quantum theory for electromagnetic radiation previously discussed in the rst four chapters,
this chapter discusses a new way of looking at an atom, in particular the introduction of
quantization concept to atomic scales yet in its primitive form as only a single quantum
number involved in describing the dynamics of the atom. To start with, we here give a brief
introduction to the development of atomic model.
The early modern, if it were worth to say, atomic model was the Thomson model in 1898,
in which an atom was viewed as a homogeneous solid sphere of uniformly distributed mass
having positive charge and negatively charged electrons distributed over the sphere surface to
produce a totally, electrically neutral atom. But, this model failed to explain the various lines
of atomic spectra found from direct observations. The mysterious lines and the corresponding
revised model were then suggested by Rutherford in 1912 after a series of experiments using
a beam of -particles scattered by a thin metal foil (see Figure 5.1). The important result of
these experiments was that most of atomic mass and its corresponding positive charge lie in
a central region of the atom called the nucleus, while the electrons surrounding it. However,
the idea of such a nuclear atom raised some fundamental questions as follows: (1) If there are
only Z protons within the nucleus comprising roughly a half of the total mass of the nucleus,
what composes the other half ? (2) What provides the cohesive force to keep protons within
a conned region of 10
14
m, the size of the nucleus ? (3) What drives the electrons to move
around the nucleus ? (4) How does this motion account for the observed spectral lines ?
Among these, Rutherford gave his best speculation to answer to the rst two questions. But,
he had certainly no answer to the other two until the work of Bohr came into play in 1913.
39
40 5. THE BOHR MODEL OF THE ATOM
Figure 5.1: A beam of -particles scattered by a dense positively charged nucleus
(taken from Ch.4, Modern Physics, Serway et al., 2005).
5.2 Bohrs Atomic Model
The Rutherford atomic model is based on the fact that the electrons must revolve the nucleus.
Meanwhile, classical electromagnetic theory of radiation claims that accelerated electrons
revolving the nucleus will radiate some energy that makes orbital radius of the electrons
become less and less. In this sense, the radius decreases steadily followed by an ever-increasing
radiant frequency corresponding to the energy released. The atom may therefore be unstable
and collapsing as the electrons plunge into the nucleus, as shown in Figure 5.2.
The wrong deductions above that lead to a continuous emission spectrum were well tackled
by Niels Bohr, a young Danish physicist. He argued that classical electromagnetic radiation,
theoretically proposed by Maxwell and experimentally conrmed by Hertz long before Bohrs
new model of the atom shook the world of physics to its foundation, is false when applied to
systems at a microscopic scale. Based on the great work of Planck, he applies the concept of
quantized levels of energy and angular momentum to orbital electrons that occupy stable
stationary states in which no energy is lost. At the same time, based on the corresponding
work of Einstein, he applies the concept of photon emitted when there is a quantum jump
of an electron from a particular stationary state associated with certain values of energy and
angular momentum to another. In this way, Bohr developed a new theory of an atom in 1913
by combining basic physic principles of classical mechanics with a simple quantum theory of
light emission. The Bohr model is then able to explain the existence of atomic spectral lines
of the hydrogen atom, in good agreement with experimental evidence and thereby resolving
the shortcomings of the previous, classical atomic models.
5.2. Bohrs Atomic Model 41
Figure 5.2: Classical radiation theory of an orbital electron having less and less energy
as it approaches the nucleus (taken from Ch.4, Modern Physics, Serway et al., 2005).
In this context, let us now examine in detail the semi-classical, Bohr theory for an atom.
As it originally applies to a hydrogen atom with a single proton and a single electron only,
the basic idea of the Bohrs theory is that the dynamics of the hydrogen is determined by
the Coulomb attraction of a nucleus (the proton) and the electron that drives the electron to
orbit the nucleus. The orbiting electron has an energy derived from a classical expression for
the total energy. This concept is then combined with two quantum postulates as follows :
The hydrogens electron is assumed to orbit the nucleus in concentric circular paths
associated with stationary orbits, each having its own angular momentum L equal to
an integer multiplied by h/2, where h is the Plancks constant. These orbits are then
referred to atomic shells. This can be mathematically expressed as
_
Ld = nh
L = n
(5.1)
where n = 1, 2, 3 .... is an integer, the so-called principal quantum number associated
with the nth stationary orbit at which the electron may be located. The stationary orbits
are also related to discrete values of energy representing quantum states, each having its
own quantum number n, for the orbiting electron. In such stable orbits, this electron
radiates no energy.
When an electron from a quantum state at a particular level of energy jumps into another
quantum state at a lower level of energy, then photon with a certain value of frequency
is emitted (see Figure 5.3). In this case, the radiant frequency is then proportional to
the dierence in the level of energy E between the two states, and is written as
=
E
h
=
E
i
E
f
h
(5.2)
42 5. THE BOHR MODEL OF THE ATOM
Figure 5.3: A photon of frequency is emitted as a result of quantum transition from
an initial state with a particular level of energy E
i
to a nal state with a lower level of
energy E
f
(taken from http://www.kottan-labs.bgsu.edu).
where E
i
and E
f
are levels of energy for the initial and nal quantum states, respectively.
Note that the other way around of atomic transition may occur when an electron absorbs
some amount of energy to move from an initial state of one energy level to a nal state
of a higher energy level. For this case, (5.2) still holds with is the absorbed frequency.
5.3 Bohrs Explanation to the Model
This section describes the uniqueness of the Bohrs model in that it combines well classical and
quantum concepts. The combined ideas then lead to discretized orbit radii and levels of the
electrons energy, and are at the same time prove the Bohrs postulates to be inter-connected.
Let us start with the hydrogen having a single electron only. From the classical point of view,
the total energy E of the hydrogens electron is given by
E =
1
2
mv
2
k
e
2
r
=
L
2
2mr
2
k
e
2
r
(5.3)
where m = 9.11 10
31
kg denotes the mass of the hydrogens electron, e = 1.6 10
19
C is
the electron charge, and k = 9 10
9
Nm
2
C
2
is the Coulomb constant for vacuum.
Next, the dynamics of such an electron can be written in the form of Newtons second law
of motion as follows,
mv
2
r
=
L
2
mr
3
= k
e
2
r
2
(5.4)
Inserting (5.1) to (5.4) yields the hydrogens electron orbit radii r
n
as follows,
r
n
=
n
2

2
mke
2
= n
2
a
o
for n = 1, 2, 3, ... (5.5)
5.3. Bohrs Explanation to the Model 43
Figure 5.4: The rst three Bohr atomic orbits for the hydrogen atom (taken from Ch.4,
Modern Physics, Serway et al., 2005).
where = h/2 and a
o
=
2
/mke
2
= 0.529

A is here called the Bohr radius. It is then clear
from (5.5) that only orbits illustrated in Figure 5.4 with certain values of quantum number n
are allowed to occupy. These discrete orbits are due to the non-classical requirement for the
electrons angular momentum to be an integral multiple of . Such orbits form atomic shells
surrounding the nucleus, where the smallest orbit is called K shell having radius r
1
= a
o
and
is associated with n = 1. For larger orbits with n > 1, the shells are called L, M, N ........
Later on, according to quantum mechanics these orbits are divided into sub-shells.
The above quantization of the electrons orbit radii for the hydrogen atom immediately
leads to the discretized levels of the electrons total energy, which will be derived below.
Substituting (5.4) into (5.3) results in
E =
L
2
2mr
2

L
2
mr
2
=
L
2
2mr
2
(5.6)
With the help of (5.1) and (5.5) to be inserted into (5.6), we can write the allowed levels E
n
of the total energy for the hydrogens electron as
E
n
=
mk
2
e
4
2
2
n
2
=
R
n
2
for n = 1, 2, 3, ... (5.7)
where R = mk
2
e
4
/2
2
= 13.6 eV is called the Rydberg constant for the electrons energy.
In other literatures, the Rydberg constant is dierently valued as it is meant for a dierent
quantity (see further 5.4). The negative sign for the electrons total energy in (5.7) suggests
that the electron is bound to the nucleus and thereby requiring some amount of energy to
44 5. THE BOHR MODEL OF THE ATOM
Figure 5.5: A diagram showing the energy-discretized states of the hydrogen atom
and some spectral lines following atomic transitions (taken from Ch.4, Modern Physics,
Serway et al., 2005).
liberate it from the Coulomb attraction. The lowest possible stationary state corresponding
to quantum number n = 1 is generally called the ground state having the smallest energy of
E
1
= 13.6 eV. The next state is the rst excited state, corresponding to quantum number
n = 2 and hence having an energy of E
2
= 3.4 eV.
An energy-level diagram for the hydrogen atom showing the energy-discretized states with
the corresponding quantum numbers is depicted in Figure 5.5. The uppermost level of energy
corresponds to n = and hence E
n
= 0, representing a state in which an electron is free
from Coulomb nuclear eld. When this electron experiences an external force it has a kinetic
energy only for which it is usually termed as a free electron. The minimum energy required to
remove the bound electron from its ground state to a large distance from the nuclear inuence
is called the ionization energy. Thus for the hydrogen, the ionization energy is 13.6 eV.
5.4 Hydrogen Spectral Lines
The great success of the Bohr atomic model for the hydrogen is achieved when it is used
to describe the existence of the hydrogen spectral lines. Instead of a continuous emission
spectrum argued by Rutherfords model, the Bohrs theory predicts that atomic transitions
between levels of energy produce a spectral series of separated lines (see Figure 5.5), each
having its own characteristics; all the observed lines in Lymann series end at a state of n = 1,
lines of Balmer series are down to n = 2, and Paschen lines towards n = 3.
5.4. Hydrogen Spectral Lines 45
Here we derive a mathematical expression for the observed hydrogen spectral lines using
(5.2) and (5.7). Let us suppose a quantum transition from an outer orbit at an initial state of
quantum number n
i
having energy E
i
= R/n
2
i
to an inner orbit at a nal state of quantum
number n
f
having energy E
f
= R/n
2
f
. Such a transition emits a photon of frequency
given by
=
R
h
_
1
n
2
f

1
n
2
i
_
(5.8)
The above equation is equivalent to
1

=
R
hc
_
1
n
2
f

1
n
2
i
_
= R
_
1
n
2
f

1
n
2
i
_
(5.9)
where represents the photon wavelength corresponding to the emitted frequency in (5.8)
and R is exactly the same as R/hc = 1.097 10
7
m
1
, dened to be the Rydberg constant
for the emitted wavelengths of the hydrogen spectral series depicted in Figure 5.5.
When Bohr proposed his theory in 1913, the spectral lines of Balmer series (n
f
= 2 and
n
i
> 2) and Paschen series (n
f
= 3 and n
i
> 3) had already been found two years before.
The Rydberg constant R in (5.9) theoretically derived from the Bohrs quantum theory for
the hydrogen is found to be in good agreement with an empirical value for a constant used to
describe Balmer and Paschen lines. This was followed by the Lymann series (1916) for which
n
f
= 1 and n
i
> 1. In the years to come since the rst three spectral series, Brackett (1922)
and Pfund (1924) observed the spectral lines for n
f
= 4 and n
f
= 5, respectively. Among all
these series, only Balmer lines constitute a range of wavelengths in a visible light spectrum.
Impressed by his successful explanation to the experimental observations of the hydrogen
spectral lines, Bohr then immediately extended his theory to hydrogen-like atoms, such
as He
+
and Li
2+
, in which all but one electron had been removed from a nucleus of positive
charge Ze, where Z denotes the number of proton. For a single electron of negative charge e
orbiting the nucleus of a hydrogen-like atom, the quantized orbit radii in (5.5) becomes
r
n
=
n
2

2
mkZe
2
= n
2
a
o
Z
for n = 1, 2, 3, ... (5.10)
and the corresponding discrete energies in (5.7) becomes
E
n
=
mk
2
Z
2
e
4
2
2
n
2
= Z
2
R
n
2
for n = 1, 2, 3, ... (5.11)
The extended theory was used to explain mysterious lines observed in hot stellar atmospheres.
In the rst paper of his great trilogy published in late 1913, Bohr noted that a formula similar
to (5.9) could account for the Pickering series (1896), a series of lines describing the spectra
46 5. THE BOHR MODEL OF THE ATOM
of stars. Bohr argued that singly ionized helium atoms would have exactly the same spectrum
as that of the hydrogen, but the corresponding wavelengths would be four times shorter.
In his eort to explain such spectral lines, Bohr even went further to take into account the
contributions of the motion of both the electron and the nucleus about their centre of mass
to derive the so-called reduced mass. Using the reduced mass of the hydrogen and He
+
, Bohr
found that the ratio of the Rydberg constants for ionized helium and hydrogen is 4.00160,
compared to the value of 4.00163 obtained from laboratory experiments of Fowler in 1912. He
then predicted the lines to be due to the presence of singly ionized helium atoms, instead of
the hydrogen.
Although Bohrs remarkable achievement in explaining both emission and absorption lines
of the hydrogen and hydrogen-like atoms and in describing the shell structure of an atom
is of paramount importance in the development of atomic model using an approach of an
uneasy mixture of classical and quantum ideas, the Bohr model of the atom is fundamentally
incomplete. The model has, however, some diculties regarding with discrete circular orbits.
These orbits give a classical, deterministic property to the position of an orbiting electron
in that the electrons position at any time can thus be determined with a high accuracy. This
property is one in which it is not acceptable in the context of quantum mechanical model of
an atom, where the electrons exact position cannot be precisely determined but only with
some condence by a probabilistic value.
The above argument is in line with Heisenbergs uncertainty principle (which will be
discussed in Chapter 7), where it is argued that the electrons of a multi-electron atom occupy
a region like a cloud around the nucleus with the cloud density being the probability to nd
an electron in a particular region. As more electrons are introduced to a stable multi-electron
atom, the paths through which the electrons move around the nucleus are complicated, and are
likely to be overlapping in elliptical orbits, making the circular orbits are no longer applicable.
This implies that additional quantum numbers other than the principal quantum number n
are needed to describe the complex structure of a multi-electron atom. Detailed calculations
of quantized energy levels of such an atom reveal the presence of atomic sub-shells, associated
with orbital quantum numbers. The complete description of all quantum numbers required to
describe the dynamics of a system at atomic scales is given in the course of Quantum Physics.
5.5 Bohrs Correspondence Principle
Bohrs correspondence principle is a simple principle that provides a smooth and gradual
change of the new but in some sense yet primitive quantum theory into classical theory
in the limit of a very large quantum number previously dened. This principle naturally comes
from the basic idea that a new theory should be able to capture all the essence of the old law
5.5. Bohrs Correspondence Principle 47
Figure 5.6: A sketch showing the classical limit of the Bohr theory as the principal
quantum number approaches innity, associated with very large orbit radii (taken from
Ch.4, Modern Physics, Serway et al., 2005).
with no exception, and at the same time to demonstrate when the new theory approaches the
old one. If this requirement is satised, then the new theory is said to be rmly established
with no doubt. A similar situation can also be found in the context of special relativity theory.
This theory shows its strength not only by combining spatial and temporal components of
the four-dimensional space-time coordinates with the concept of the four-vector previously
discussed in Chapter 1, but also by demonstrating that calculations of the physical quantities
based on the special relativity gradually change into classical results of Newtonian mechanics
when the speed is relatively small compared to the speed of light.
In the hands of Bohr, the correspondence principle becomes a master tool to demonstrate
where the quantization of the electrons orbital angular momentum comes from. Once derived,
the validity of the quantization concept depends upon the agreement with measurements of
lines of atomic spectra. Here, we simply use Figure 5.6 to derive the rst Bohr postulate
earlier mentioned in 5.2 from which the quantized electrons orbit radii shown in (5.5) and
the discretized electrons total energy in (5.7) are derived. Let us assume that r
1
and r
2
are
the radii of the two adjacent quantum orbits, each having its own orbital angular frequencies

1
and
2
, respectively. We go further by assuming that r
1
r
2
r and hereby
1

2
.
We also assume that

is the frequency of a photon emitted as a result of a quantum transition


from the initial state of radius r
1
to the nal state of radius r
2
.
Using the fact that the orbiting electron is under the inuence of the Coulomb force, we
can show that the total energy for the electron is given by
E =
mk
2
e
4
2L
2
(5.12)
48 5. THE BOHR MODEL OF THE ATOM
Note that we obtain the close relationship of the electrons energy and the angular momentum
for the Bohr model. Taking the derivative of (5.12), we have
dE
dL
=
mk
2
e
4
L
3
=
mk
2
e
4
(mk
2
e
4
)/
= (5.13)
or it can be simply written as
dE = dL (5.14)
where dE is the energy of a photon emitted when a quantum transition from r
1
to r
2
occurs
and dL is the corresponding change in the electrons angular momentum. Thus based on
Figure 5.6 and the denitions of dE and

, we have dE =

and so (5.14) becomes


= dL (5.15)
As earlier noted, the correspondence principle must be able to predict the same frequency
occurs for the emitted photon following the transition, as suggested by the quantum theory of
light emission, and for the electrically charged electron moving around the nucleus, based on
classical electromagnetic theory of radiation. In this way, it is true that

= , and therefore
we can write
dL = (5.16)
It is understood from (5.16) that the electrons orbital angular momentum in a specic orbit
is an integral multiple of , from which the rst postulate of Bohr L = n given in (5.1) is
proposed. What an amazing result ! Bohr realized that although (5.16) originally was derived
for the special case of orbits with large radii, it was a universal quantum principle that it was
relevant to a wider range of applicabilities than the classical Maxwells theory of radiation.
5.6 Exercises
1. If an electron moves from an inner orbit to an outer orbit, does its total energy increase
or decrease ? Does its kinetic energy increase or decrease ?
(taken from Ch.37, Physics for scientists and engineers, Tipler, 1999).
2. Show that the speed of an electron in the nth Bohr orbit of the hydrogen atom is given
by v
n
= e
2
/2
o
nh.
(taken from Ch.37, Physics for scientists and engineers, Tipler, 1999).
3. The binding energy of an electron is the minimum energy required to remove the electron
from its ground state to a large distance from the nucleus.
5.6. Exercises 49
(a) What is the binding energy for the hydrogen atom ?
(b) What is the binding energy for He
+
?
(c) What is the binding energy for Li
2+
?
Note that a singly ionized helium He
+
and a doubly ionized lithium Li
2+
behave as
hydrogen-like atoms in that such ionized elements consist of a positively charged nucleus
and a single bound electron.
(taken from Ch.37, Physics for scientists and engineers, Tipler, 1999).
4. What is the radius of the rst Bohr orbit in He
+
and Li
2+
?
(taken from Ch.4, Modern Physics, Serway et al., 2005).
5. A hydrogen atom is in its ground state. Using the Bohr theory of the atom, calculate
(a) the radius of the orbit
(b) the linear momentum
(c) the angular momentum
(d) the kinetic energy
(e) the potential energy
(f) the total energy of the electron
(taken from Ch.4, Modern Physics, Serway et al., 2005).
6. A photon is emitted when a hydrogen atom undergoes a jump of electronic transition
from the initial quantum state of n = 3 to the nal quantum state of n = 2. Determine
(a) the energy
(b) the wavelength
(c) the frequency of the emitted photon
(taken from Ch.4, Modern Physics, Serway et al., 2005).
7. A hydrogen atom initially at rest in the n = 3 quantum state decays to the ground state
with the emission of a photon.
(a) Calculate the wavelength of the emitted photon
(b) Estimate the recoil momentum of the atom
(c) Estimate the kinetic energy of the recoiling atom
(d) Where does this energy come from ?
(taken from Ch.4, Modern Physics, Serway et al., 2005).
50 5. THE BOHR MODEL OF THE ATOM
8. If one assumes that in a stationary state of the hydrogen atom the electron ts into a
circular orbit with an integral number of wavelengths, one can reproduce the results of
the Bohr theory. Work this out.
(taken from Ch.1, Quantum Physics, Stephen Gasiorowicz, 1996).
9. Use the Bohr quantisation rules to calculate the energy levels for a harmonic oscillator,
for which the energy is p
2
/2m + m
2
r
2
/2 directly given by the driving force m
2
r.
Restrict yourself to circular orbits. What is the analog of the Rydberg formula ? Show
that the corresponding principle is satised for all values of the principal quantum
number n used in quantizing the angular momentum.
(taken from Ch.1, Quantum Physics, Stephen Gasiorowicz, 1996).
10. A muon is a particle with a charge equal to that of electron and a mass equal to 207
times the mass of an electron. Muonic lead is formed when Pb
208
captures a muon to
replace an electron. Assume that the muon moves in such a small orbit that it sees a
nuclear charge of Z = 82. According to the Bohr theory, what are the radius and energy
of the ground state of muonic lead ? Use the concept of reduced mass here.
(taken from Ch.4, Modern Physics, Serway et al., 2005).
Chapter 6
THE WAVE PROPERTIES OF
SUB-ATOMIC PARTICLES
As widely known, classical wave theory suggests that the propagation of light is considered
as a natural wave phenomenon. Examples of this can be found in many physical situations.
In particular, the laws of geometric optics are empirically proved to be held when light is
incident on a surface (reection) or arrives at a boundary between two media (refraction).
In other optical phenomena, such as interference and diraction of light, the laws of physical
optics are observed to occur when a beam of light rays passes through a very small aperture.
In Chapter 4, however, it was shown that electromagnetic radiation in the form of photons
may behave as particles when interacting with matter, such as those in the photo-electric and
Compton eects. A somewhat bizarre, fundamental question based on the reverse mechanism
is raised: can a classical particle, say an electron, with its own character behave as a wave
with a totally dierent character ? If this is the case, then the revolutionary idea about
matter waves that leads to the new concept of the way we are looking at the dynamics of
moving particles at microscopic scales has to be proven by both theoretical considerations and
experimental measurements. These topics are the primary issues we discuss in this chapter.
6.1 De Broglie Hypothesis
In classical physics, the concept of particle distinguishes both qualitatively and quantitatively
from that of wave. In the context of a classical particle, it is common to characterize particle
with its inertial mass; a particle is spatially localized at a certain time. Whereas a wave is
characterized by its spatial periodicity, i.e., wavelength, or temporal periodicity i.e., frequency
for which it is also dened as an energy propagated from one point to another in space-time
coordinates; the wave spreads and hence its energy distributes over space. It is clear from this
that none can demonstrate itself to be a particle and at the same time behaves as a wave.
51
52 6. THE WAVE PROPERTIES OF SUB-ATOMIC PARTICLES
Figure 6.1: The equivalence between (a) a moving particle of mass m and speed v
o
with (b) a wave packet of wavelength and speed v
g
(taken from Ch.5, Modern Physics,
Serway et al., 2005).
In 1925, de Broglie postulated a radical idea with no experimental supports at the time
it was proposed. His postulate was nally known as the de Broglie hypothesis derived
from theoretical considerations of classical wave, relativity theory, and quantum concept for
a photon with zero rest mass. The mathematical expression of his postulate is as follows,
=
h
p
(6.1)
where h is the Plancks constant, and p are the wavelength and momentum of the photon,
respectively. At this stage, de Broglie argued that the above expression holds also for any
moving object of mass m. The following paragraphs show logical reasons for this.
De Broglie generalized (6.1) to introduce the concept of a matter wave by assuming that
a moving particle can be in principle viewed as a wave packet or a wave group. A wave
of this type must reect the fact that such a particle has a large probability of being found
within a small, conned region of space at a limited time (see Figure 6.1). It follows that a
single traveling sinusoidal wave with constant amplitude and innite extent is not relevant to
model the particle. Instead, a group of waves of limited spatial extent consisting of individual
waves with dierent wavelengths can then represent the particle. In this case, the resulting
wave group travels at a speed v
g
dened to be the group velocity of the wave, which is
identical to the observed speed v
o
of the corresponding particle.
Let us then apply the de Broglie hypothesis given in (6.1) to a photon of energy E =
and momentum p = k. These quantities are attributed to a wave character through two
related parameters, namely the angular frequency = 2 and the wave number k = 2/.
6.1. De Broglie Hypothesis 53
From these quantities, or parameters, we have = ck for the photon. If the group velocity of
de Broglie wave corresponding to this photon is dened as
v
g
=
d
dk
=
dE
dp
(6.2)
then we have v
g
= c for the speed of the wave group, as it should be for photons with zero
rest mass, m
o
= 0.
As noted, we have already made (6.1) true for photons. An interesting question is that
whether such an equation also holds for a classical particle of mass m moving at speed v.
Here, we provide a somewhat crude argument for this. During its motion, the particle has
momentum p = mv and kinetic energy E = p
2
/2m. If the motion is considered as a wave
group, then the group velocity can be calculated by assuming d/dk to be equivalent to
dE/dp and by inserting this equivalence into the denition of v
g
to get the result for v
g
= v,
consistent with the particles speed. The result also implies that, for photons, the wave group
travels at speed c. Based on this simple calculation, it is shown that (6.1), originally derived
for photons, is relevant to any moving particle with non zero rest mass.
A more detailed use of (6.1) for any moving particle with non zero rest mass is given here,
as we are now in a position to apply the de Broglie hypothesis to sub-atomic particles, such
as a proton, a neutron, and an electron. In doing so, we need to dene what is called the
phase speed, that is, the speed of a point of constant phase on a wave. The phase speed of
the wave is given by
v
p
=

k
=
E
p
(6.3)
where E dan p represent the relativistic expressions for the total energy and momentum of the
particle, respectively. We can also write the phase speed in (6.3) as a function of k only with
the help of E
2
= E
2
o
+ (pc)
2
, where E
o
is the particles rest energy. Inserting this relativistic
relation into (6.3) results in
v
p
= c

1 +
_
mc
k
_
2
(6.4)
where m is the particles rest mass. As k varies with wavelength, or equivalently frequency,
then (6.4) is meant for individual waves of the wave packet. Note that the expression in (6.4)
is also called a dispersion relation for the phase speed of each component of the wave group.
The group velocity of the wave group can then be calculated from
v
g
= v
p
+ k
dv
p
dk
(6.5)
to be evaluated at k
o
, the central wave number of a continuous distribution of wavelengths
constituting the wave group.
54 6. THE WAVE PROPERTIES OF SUB-ATOMIC PARTICLES
Substituting (6.4) to (6.5) and after some simple algebra, we can write the group velocity
of the de Broglie wave as
v
g
=
c
_
1 + (mc/k)
2
=
c
2
v
p
(6.6)
Let us go back for a moment to the phase speed given in (6.3). Solving for this phase speed
yields
v
p
=
E
p
=
mc
2
mv
=
c
2
v
(6.7)
where is previously dened in (1.3) in Chapter 1 and v is the particles speed. We eventually
get the right expression for the group velocity by inserting (6.7) into (6.6) to have v
g
= v.
The nal result obtained for v
g
is convincing in that the group velocity of the wave packet is
the same as the corresponding particles speed, as expected.
6.2 Implications of the De Broglie Hypothesis
Although it looks so simple, the de Broglie hypothesis dened in (6.1) is actually powerful.
The hypothesis combines well fundamental aspects of both classical wave of wavelength
and classical particle of momentum p. The Plancks constant h serves as a connecting bridge
between the two properties, implying that the application of the hypothesis is limited only
to the case of moving particles at a microscopic scale. In this context, the hypothesis is then
used to validate the particle property of a photon and the quantization of the orbital angular
momentum of the electron for the Bohr model of the hydrogen atom. These two implications
will be discussed in the following paragraphs.
It has been known that photons have zero mass, will always be moving at speed c, and may
behave as particles. This particle property, however, would have made photons to have mass.
This seems to be contradictory with the fact that photons are massless. To examine this, let
us dene what is called the eective inertial mass of a photon, a quantity describing how
a photon responds to an applied force acting on it. The photons eective inertial mass m
e
may reasonably be taken to be proportional to its total relativistic energy E as follows,
m
e
=
E
c
2
=
h
c
2
=
h
c
(6.8)
for which (6.8) can be rearranged to become
=
h
m
e
c
=
h
p
e
(6.9)
which is identical to (6.1) in the sense that the associated wavelength of a moving particle is
inversely proportional to its momentum, with m
e
c being the photons eective momentum.
6.2. Implications of the De Broglie Hypothesis 55
Figure 6.2: De Broglie standing waves in Bohrs stationary orbits (taken from Ch.4,
Konsep Fisika Modern, The Houw Liong, 1987, adapted from the work of Arthur Besier).
Note that by (6.8) and (6.9) the de Broglie hypothesis demonstrates its-self to be self consistent
with both relativity and quantum theories.
Another important aspect of the de Broglie hypothesis dened in (6.1) is that it provides
a physical feature of the Bohrs quantum theory for the atoms. Although the Bohrs model
is useful in describing the dynamics of the hydrogen and hydrogen-like atoms, it has also
some shortcomings regarding with, say for example, only certain values of electronic energy
are allowed to occupy in the model. De Broglie recognized this problem and made it clear
by visualizing the orbiting electrons around the nucleus as standing waves bent into circles of
discrete radii (see Figure 6.2). This point of view actually comes from the simple but brilliant
idea that an atom connes its orbiting electrons to a very small atomic dimension, in which
case the wave nature of the electrons should predominate over their particle property. From
classical wave theory we know that when a wave is restricted to a small area, then only a
discrete set of standing waves are possible to occur within that area. The Bohrs stationary
orbits are thus viewed as a result of constructive interference of these waves for a range of
wavelengths. This constructive interference ts into the circumference of circular orbits and
56 6. THE WAVE PROPERTIES OF SUB-ATOMIC PARTICLES
corresponds to an integral number of the de Broglie waves, each having wavelength . We can
then express this de Broglies argument by rewriting the rst Bohrs postulate given in (5.1)
for the quantized orbital angular momentum L = n of the hydrogens electron to start from.
After a simple algebra we have
2r
n
= n for n = 1, 2, 3, ... (6.10)
where r
n
is the Bohrs orbit radii previously dened in (5.5). The fact that the de Broglie
hypothesis originating from the wave nature of the orbiting electron revolving the nucleus
matches with the Bohrs quantum theory for the particle-like property of the electron is a key
to understanding the nature of microscopic world. Following this great discovery, scientists
have started realizing that electrons and hence other sub-stomic particles have a dual property,
the so called particle-wave duality. It can then be inferred from many cases considered that
it is not possible to observe both the particle and wave properties simultaneously. Rather, one
property completes the other for which Bohr called this as a complementary principle.
6.3 The Davisson-Germer Experiment
Following his successful doctoral degree, de Broglie as a consequence of his postulate suggested
that a stream of electrons passing through a very small aperture would produce diraction
pattern, as it would for a beam of light. In this context, a series of laboratory experiments
using Davisson-Germer experimental apparatus (see Figure 6.3) were completed in 1927 to
test the de Broglie hypothesis of a matter wave. The primary result of these experiments
was that electrons experience diraction, providing convincing support for the wave nature
of an electron. The apparatus allows for the variations of three experimental parameters,
namely the energy of an electron, the orientation of a nickel target, and the angle of an
elastic scattering. Initially, the progress of these experiments seemed to give nothing. After
an unexpected but fortunate experimental accident occurred, however, Davisson and Germer
realized that further analyses on the experimental procedure give dierent results that exhibit
a diraction pattern. Thus, they took this peculiar chance to calculate the wavelength of an
electron using both the de Broglie hypothesis given in (6.1) and a simple diraction formula,
corresponding to various values of the three parameters. In their theoretical calculation on
the basis of conservation of energy, Davisson and Germer used the non-relativistic expression
for the speed v of the electron as follows,
1
2
mv
2
= P (6.11)
where m = 9.11 10
31
kg is the electrons rest mass and P is the potential energy supply.
6.4. Exercises 57
Figure 6.3: The Davisson-Germer experimental apparatus, used for examining the
wave nature of an electron by elastically scattering a beam of low-speed electrons from
a polycrystalline nickel target (taken from Ch.5, Modern Physics, Serway et al., 2005).
The remaining step is straight forward, solving for v from (6.11) and inserting the result
into (6.1) yields
=
h

2meV
(6.12)
where e = 1.6 10
19
C is the electron charge and V is the experimental voltage supply. For
the special case of corresponding to the diraction maximum where V = 54 volt, we have
= 1.67 10
10
m, in good agreement with = 1.65 10
10
m obtained from formula for
constructive interference,
d sin = n (6.13)
where d = 2.15 10
10
m obtained from previous measurements of X-ray diraction, = 50

and n = 1 corresponds to the diraction maximum pattern.


6.4 Exercises
1. Is light a wave or a particle ? Is an electron a particle or a wave ? Support your answer
by citing specic experimental evidence.
(taken from Ch.5, Modern Physics, Serway et al., 2005).
58 6. THE WAVE PROPERTIES OF SUB-ATOMIC PARTICLES
2. If matter has a wave nature, why is this wave-like character not observable in our daily
experiences ?
(taken from Ch.5, Modern Physics, Serway et al., 2005).
3. Show that the de Broglie wavelength of an electron of kinetic energy E, measured in eV,
is given by

e
=
12.3 10
10
E
1/2
m
and that of a proton is given by

p
=
0.29 10
10
E
1/2
m
(taken from Ch.2, Introductory Quantum Mechanics, Richard Libo, 1980).
4. Show that in order to associate a de Broglie wavelength with the propagation of photons
(electromagnetic radiation), photons must travel with the speed of light c and their rest
mass must be zero. Do this relativistically.
(taken from Ch.2, Introductory Quantum Mechanics, Richard Libo, 1980).
5. Calculate the de Broglie wavelength for
(a) an electron of kinetic energy 250 eV (the rest energy of the electron is 0.511 MeV)
(b) a neutron of kinetic energy 0.02 eV (the rest energy of the neutron is 940 MeV)
(c) a proton of kinetic energy 2 MeV (the rest energy of the proton is 938 MeV)
(taken from Ch.17, Physics for scientists and engineers, Tipler, 1999).
6. To observe small objects, one measures the diraction of particles whose de Broglie
wavelength is approximately equal to the objects size. Find the kinetic energy, measured
in eV, required for electrons to resolve
(a) a large organic molecule of size 10 nm
(b) atomic features of size 0.10 nm
(c) a nucleus of size 10 fm
Repeat these calculations using alpha particles in place of electrons.
(taken from Ch.5, Modern Physics, Serway et al., 2005).
7. Find the de Broglie wavelength of a ball of mass 0.2 kg just before it strikes the Earth
after being dropped from a building 50 m tall.
(taken from Ch.5, Modern Physics, Serway et al., 2005).
6.4. Exercises 59
8. An electron has a de Broglie wavelength equal to the diameter of the hydrogen atom.
What is the kinetic energy of the electron ? How does this kinetic energy compare with
the ground-state energy of the hydrogen atom ?
(taken from Ch.5, Modern Physics, Serway et al., 2005).
9. For an electron to be conned to a nucleus, its de Broglie wavelength would have to be
less than 10
14
m.
(a) What would be the kinetic energy of an electron conned to this region ?
(b) On the basis of this result, would you expect to nd an electron in a nucleus ?
(taken from Ch.5, Modern Physics, Serway et al., 2005).
10. The dispersion relation for free electron waves is
(k) =
_
c
2
k
2
+ (mc
2
/)
2
From the above equation, obtain expressions for the phase velocity v
p
and the group
velocity v
g
of these waves and show that their product is constant, independent of k.
From your results, what can you learn about v
g
when v
p
> c ?
(taken from Ch.5, Modern Physics, Serway et al., 2005).
60 6. THE WAVE PROPERTIES OF SUB-ATOMIC PARTICLES
Chapter 7
THE HEISENBERG UNCERTAINTY
PRINCIPLE
One of the most basic and far-reaching concepts of modern physics is Uncertainty Principle,
proposed by Werner Heisenberg in 1927. This beautiful principle is fundamentally concerned
with the limit of our ability in simultaneously determining with high accuracy and precision
two independent but conjugate variables in physics, which can be in the form of a pair of two
quantities, such as position and momentum, energy and time, angle and angular momentum.
The limit itself has nothing to do with imperfections in practical measuring instruments, or
equivalently with careless measurements. Rather, it arises automatically from the nature of
microscopic systems. Although this principle in its simple form is later known to be one of
cornerstones in modern physics, Heisenberg is also famous for his contributions to develop a
complete theory of quantum mechanics. Indeed, modern quantum theory was pioneered
by the works of Schrodinger and Heisenberg, and some other physicists in the late of 1920s.
Apparently, there were two dierent quantum theories, namely wave mechanics proposed
by Schr odinger and matrix mechanics suggested by Heisenberg. Although the latter is quite
elegant, it only paid little attention to physicists at the time it was launched for some reasons.
The matrix mechanics involved an unfamiliar way of complicated mathematics to describe the
dynamics of sub-atomic particles, and was arguably based on rather vague physical concepts.
However, these two formulations were later shown to be completely equivalent, in the sense
that both theoretical approaches conrm that the principal quantum number is not adequate
to describe the dynamics of microscopic particles. Instead, additional parameters associated
with orbital, magnetic, and intrinsic quantum numbers are needed to complete an atomic
description. A detailed derivation of these quantum numbers will be given in the course of
Quantum Physics. Instead, this chapter focuses on the Heisenberg uncertainty principle,
providing a mathematical basis for the principle and exploring its consequences.
61
62 7. THE HEISENBERG UNCERTAINTY PRINCIPLE
x
f(x)
-L
0
Eo
L
A(k)
0
(a)
(b)
kp -kp
k
kp+/L kp-/L
EoL
Figure 7.1: The idealised prole of (a) a cosine pulse of width 2L and amplitude E
o
,
and (b) its Fourier transform with a carrier spatial frequency k
p
in the space domain.
7.1 Mathematical Basis for the Uncertainty Principle
In the previous chapter, it was stated that what was meant by a wave packet or a wave group
associated with a moving particle was a group of waves of limited spatial extent consisting of
individual waves of dierent frequencies. Wave of this type is commonly found in the form
of a pulse propagated in space. The term space here is used in a general sense, for which it
could be either x-axis or t-axis depending upon an explicit mathematical function of the pulse.
To examine how a pulse is transformed from one form to another in a particular domain, here
we consider an idealised harmonic pulse in the form of a cosine periodical function, that is,
the cosine wavetrain in the space domain (see Figure 7.1a) given by
f(x) = E
o
cos k
p
x for L x L
= 0 otherwise
(7.1)
The choice to work in the space domain is optional in that the time-dependent disturbance
is equivalently applicable (see also Figure 7.2). In the time domain, the corresponding cosine
wavetrain is written as
g(t) = E
o
cos
p
t for T x T
= 0 otherwise
(7.2)
Here we solve the problem using spatial variables in (7.1) than using temporal variables
in (7.2), as it is easier to examine the wave prole at t = 0 rather than the prole at x = 0.
7.1. Mathematical Basis for the Uncertainty Principle 63
t
g(t)
-T
0
Eo
T
a()
0
(a)
(b)
p -p

p+/T p-/T
EoL
Figure 7.2: The idealised prole of (a) a cosine pulse of width 2T and amplitude E
o
,
and (b) its Fourier transform with a carrier temporal frequency
p
in the time domain.
As f(x) given in (7.1) is an even function, the corresponding cosine Fourier transform
A(k) in the space domain is calculated from
A(k) =
_

f(x) cos kx dx (7.3)


The above integral tells us about how dynamic variables k and x are interchangeable, meaning
that a function of a particular variable in a given domain can be transformed into another
function of its counter-part variable in the same domain. Thus based on Figure 7.1 and by
substituting (7.1) into (7.3), we have
A(k) =
_
L
L
E
o
cos k
p
x cos kx dx
=
_
L
L
E
o
2
_
cos(k
p
+ k)x + cos(k
p
k)x
_
dx
= E
o
L
_
sin(k
p
+ k)L
(k
p
+ k)L
+
sin(k
p
k)L
(k
p
k)L
_
= E
o
L
_
sinc (k
p
+ k)L + sinc (k
p
k)L

(7.4)
Several interesting possibilities are here discussed regarding with (7.4). When there are
many waves in the train, that is,
p
L, then the left and right peaks are both narrow
and widely spaced. When k = k
p
, the second sinc function reaches a maximum value of one,
meaning that A(k) has a maximum value of E
o
L (Figure 7.1b). In particular, the rst sinc
64 7. THE HEISENBERG UNCERTAINTY PRINCIPLE
function in the bracket becomes unimportant hence negligible compared with the second one.
Thus, the Fourier transform of the train becomes
A(k) = E
0
Lsinc (k
p
k)L (7.5)
which is the right-hand side curve of Figure 7.1b. In the case of the time-dependent cosine
wavetrain, the corresponding Fourier transform is given by
a() = E
0
T sinc (
p
)T (7.6)
which is the right-hand side curve of Figure 7.2b. Here, k and are correlated to one another
by the phase velocity. As the train becomes innitely long, that is, L or accordingly
T , the curve closes down to a single tall spike at k
p
or again accordingly
p
.
Notice that the square of A(k) or a() represents the energy density of the wave packet.
Most of this energy is carried away by the train, associated with part of the curve with a
spatial frequency k ranging from k
p
/L to k
p
+/L (see Figure 7.1b). A longer train leads
to the energy to be concentrated in a narrower range of k about k
p
. If the train were innitely
long, there would be singularity in the Fourier transform, in sense that all the energy would
be focused on a single value of frequency at k
p
or correspondingly
p
. This is the perfectly
limiting case of the idealised monochromatic wave.
In real cases, however, there have been no innitely long wavetrain such that no singularity,
hence no single value of frequency in the Fourier transform. Thus, it is sensible to calculate the
width of either spatial frequency k or temporal frequency , in which range the energy of
a given pulse is transferred. This is clearly illustrated by the right curves in Figures 7.1b and
7.2b. For the space-dependent pulse, the spatial frequency bandwidth k is calculated
from
k = k
2
k
1
=
_
k
p
+

L
_

_
k
p


L
_
=
2
L
(7.7)
and equivalently for the time-dependent pulse, the temporal frequency bandwidth is
given by
=
2

1
=
_

p
+

T
_

_


T
_
=
2
T
(7.8)
It is also clear from Figures 7.1a and 7.2a that the spatial extent x or temporal extent t
of the pulse is xed at a value of x = 2L or t = 2T, respectively. The product of these
frequency widths and their extents of the associated pulses are therefore given by
k x =
2
L
2L = 4
t =
2
T
2T = 4
(7.9)
7.2. Interpretations of the Uncertainty Principle 65
It follows that if the wave packet has a narrow frequency, spatial or temporal, bandwidth then
the packet will spread out in a large region of space or time. In quantum mechanics, a wave
packet or group represents a moving microscopic particle having nite energy and momentum.
If the narrow frequency bandwidth is associated with a denite value of energy or momentum,
then the exact location of the particle in space-time coordinates becomes highly uncertain.
Note that the precise value of the number on the RHS of (7.9) relies upon the functional
form of f(x) of the pulse, the geometrical shape chosen for A(k), and on the specic denitions
of k and x. A dierent choice of f(x) or A(k) and/or a dierent rule for dening k and
x will give a slightly dierent number. For the case of a Gaussian wave packet and use of
standard deviations for the denitions of both k and x, or correspondingly and t,
we have
k x
1
2
and t
1
2
(7.10)
for which it can be easily shown that
p x

2
and E t

2
(7.11)
In words, the uncertainty relations above state that the more precisely the momentum (energy)
is determined the less precisely the position (time) is known, and conversely. Here at this stage,
two equivalent questions are addressed : what do the uncertainty relations tell us, anyway ?
and how do we interpret them ? These issues will be discussed in 7.2.
But before doing so, it is better here to provide a three-dimensional case for the rst
formulation of the uncertainty principle in (7.11) by rewriting it as
p
x
x

2
p
y
y

2
p
z
z

2
(7.12)
While the momentum-position uncertainty relation given in (7.12) is relatively clear,
the energy-time uncertainty relation in the second formulation of (7.11) conrms that
the precision with which we are able to determine accurately the energy of a given system is
limited by the time taken for measuring the energy of such a system.
7.2 Interpretations of the Uncertainty Principle
As earlier mentioned, the term uncertainty may refer to either the uncertainty resulting from
a series of experimental measurements, meaning that it is an error in the measurements, or
the uncertainty derived from conceptual framework. The dierence in the meaning between
the two is that the uncertainty in the latter case arises from the product of uncertainties of
dierent variables (not from the uncertainty in any single quantity). Let us suppose here
66 7. THE HEISENBERG UNCERTAINTY PRINCIPLE
that the probability distributions for two physical quantities are not independent; they are
constrained such that both distributions cannot be made arbitrarily narrow. For example, it
could be the case that when one quantum distribution is made narrow, then the other must
be relatively wide so that the product of the two uncertainties satises some lower bound.
This is the truly essence of the quantum-mechanical uncertainty relations.
Let us here discuss further consequences of the quantum-mechanical uncertainty relations.
Take the rst formulation of (7.11) for a classical moving particle, where its momentum and
position are both independent and well dened in space-time coordinates. There is no doubt to
say that the particles position does not correspond to its momentum, and vice versa, and that
the two quantities can be specied to arbitrary precision. Therefore, it is possible to determine
precisely these quantities at any time in space once the velocity of the particle is known. This
behaviour is considered as a basic principle of determinism in classical mechanics. However,
the quantum-mechanical uncertainty relations state that quantum probability distributions
cannot be made independent. It is, for example, impossible for a microscopic particle that is
represented by a quantum state or a wavefunction to have a nite probability distribution
in position and at the same time no spread in its momentum, or conversely. What we can
only say is that the probability to nd such a particle in its particular state is given by
2
integrated over the whole space. According to quantum mechanics, it is thus meaningless to
get one dynamic variable precisely determined, as its corresponding conjugate variable must
be highly uncertain at the same time. This fundamental consequence makes the nature of
quantum measurements for microscopic world to be undeterministic, as opposed to classical
mechanics. Based on this basic physics principle, all measurements of physical quantity can
only be determined with some condence. This has further implications for the Bohrs theory,
which will be discussed below.
As previously discussed in Chapter 5, the Bohr model of the hydrogen atom leads to
discretized energy levels that agree well with spectroscopic data. Despite its usefulness, this
model has also some shortcomings for the following reasons. In fact, there is no basis theory for
justication of the well-dened stationary orbits. The nature of orbiting electrons is dicult
to predict as the exact location of such electrons cannot be precisely determined, as suggested
by the uncertainty principle. Rather, the electrons occupy a region like cloud surrounding
the nucleus, with the cloud density being the probability that these electrons could be found
in that region. As a consequence, the orbiting electrons may have overlapping states, making
the Bohrs circular orbits no longer applicable.
For concluding remarks, it is argued that the Heisenberg uncertainty relations in (7.11)
in their original forms are therefore not just about mathematical results, as they put forward
profound physical and philosophical implications. Indeed, the uncertainty relations remain an
important element in the structure of modern quantum mechanics to date.
7.3. Exercises 67
7.3 Exercises
1. Atomic nuclei, typically of size 10
14
m, frequently emit electrons, with energies of
about 1 MeV or less. Use the uncertainty principle to show that hypothetical electrons
trapped electrons within a conned region having an energy of order 20 MeV could
not be contained in the nucleus before the decay.
(taken from Ch.2, Quantum Physics, Stephen Gasiorowicz, 1996).
2. Use the Heisenberg uncertainty principle to estimate the ground state energy of a one
dimensional harmonic oscillator. The total energy for the oscillator is given by
E =
p
2
2m
+
1
2
m
2
x
2
(taken from Ch.2, Quantum Physics, Stephen Gasiorowicz, 1996).
3. Monochromatic light with wavelength of 6000

A passes through a fast shutter that
opens for only 10
9
second. What will be the spread in wavelengths in the no longer
monochromatic light ?
(taken from Ch.2, Quantum Physics, Stephen Gasiorowicz, 1996).
4. The size of an atom is approximately 10
8
cm. To locate an electron within the atom,
one should use electromagnetic radiation of wavelength not longer than, say, 10
9
cm.
(a) What is the energy of a photon with such a wavelength (in eV) ?
(b) What is the uncertainty in the electrons momentum if we are uncertain about its
position by 10
9
cm ?
(taken from Ch.2, Introductory Quantum Mechanics, Richard Libo, 1980).
5. A proton has a kinetic energy of 1 MeV. If its momentum is measured with an uncertainty
of 5%, what is the minimum uncertainty in its position ?
(taken from Ch.5, Modern Physics, Serway et al., 2005).
6. We wish to measure simultaneously the wavelength and position of a photon. Assume
that the measurement gives = 6000

A with an accuracy of one part in a million, that


is, / = 10
6
. What is the minimum uncertainty in the position of the photon ?
(taken from Ch.5, Modern Physics, Serway et al., 2005).
7. A beam of electrons is incident on a slit of variable width. If it is possible to resolve a 1%
dierence in momentum, what slit width would be necessary to resolve the interference
pattern of the electrons if their kinetic energy is
68 7. THE HEISENBERG UNCERTAINTY PRINCIPLE
(a) 0.01 MeV ?
(b) 1.0 MeV ?
(taken from Ch.5, Modern Physics, Serway et al., 2005).
8. An excited nucleus with a lifetime of 0.1 ns emits a ray of energy 2 MeV. Can the
energy width (uncertainty in energy is E) of this 2 MeV emission line be directly
measured if the best gamma detectors can measure energies to 5 eV ?
(taken from Ch.5, Modern Physics, Serway et al., 2005).
9. A
o
meson is an unstable particle produced in high energy particle collisions. It has a
mass-energy equivalent of about 135 MeV, and it exists for an average lifetime of only
8.7 10
17
second before decaying into two rays. Using the uncertainty principle,
estimate the fractional uncertainty m/m in its mass determination.
(taken from Ch.5, Modern Physics, Serway et al., 2005).
10. One of the cornerstones of quantum mechanics is that bound particles cannot be in
stationary states even at zero absolute temperature. A bound particle is one that
is conned in some nite region of space, as is an atom in a solid. There is a non-zero
lower limit on the kinetic energy of such a particle. Suppose a particle is conned in
one dimension to a region of width L. Obtain an approximate formula for its minimum
kinetic energy.
(taken from Ch.4, Modern Physics, Randy Harris, 2007).
Appendix A
QUIZ on The Relativity Theory
The following questions are all taken from Introductory Classical Mechanics, Morin, 2003.
1. Does the special-relativistic time dilation depend upon the acceleration of the moving
clock ? Argue for your answer.
2. A fast train moves at a speed of 4c/5. A clock is thrown from the back of the train to
the front. As measured from the ground reference frame, the time of ight is 1 second.
Is the following reasoning correct ? The factor between the train and the ground is
determined by = 1/
_
1 (4/5)
2
= 5/3. As moving clocks run slow, and therefore the
time elapsed on the clock during the ight is 3/5 of a second. Argue for your answer.
3. Two twins travel away from each other at relativistic speeds. The special-relativistic
time dilation eect says that each twin sees the others clock running slow, and therefore
each says the other has aged less. How would you reply to or comment on the following
someones ask : Which twin is really younger ?
4. Elementary particles called muons, which are identical to electrons, except they are
about 200 times as massive, are created in the upper atmosphere when cosmic rays
collide with air molecules. The muons have an average lifetime of about 2 10
6
s,
then they decay into electrons, neutrinos, and the like, and move at nearly the speed of
light. Assume for simplicity that a certain muon is created at a height of 50 km, moves
straight downward, has a speed of 0.99998 c, decays in exactly 2 10
6
s, and does not
collide with anything on the way down. Will such a muon reach the earth before it
decays ?
5. The momentum of an object with rest mass m and speed v is given by p = mv. Is the
following statement correct : A photon with zero mass should have zero momentum.
Argue for your answer.
69
70 A. QUIZ on The Relativity Theory
6. It is not necessary to postulate the impossibility of accelerating an object of rest mass m
to speed c. It follows as a logical consequence of the relativistic form of total energy.
Explain adequately.
7. Accepting the facts that the energy and momentum of a photon with zero rest mass are
E = h and p = h/c, derive the relativistic formulas for the energy and momentum of
a massive particle, E = mc
2
and p = mv. Hint: consider a mass m that decays into
two photons. Look at this decay in the rest frame of the mass and a frame of reference
in which the mass moves at speed v. You will need to use the Doppler eect here.
8. A plane ies at a constant height of h. What should its speed be so that an observer
on the ground sees the planes clock tick at the same rate as a ground clock ? Assume
that the speed v of the plane is much small compared to the speed of light c.
9. A clock starts on the ground and then moves up a tower at constant speed v. It sits on
the top for a time T and then descends at constant speed v. If the tower has height h,
how long should the clock sit at the top so that it comes back showing the same time
as a clock that remained on the ground ? Assume that the speed v of the plane is much
small compared to the speed of light c.
10. A person B moves at a speed of v which is much small compared to the speed of light c
in a circle of radius r around another person A. By what fraction does Bs clock run
slower than As ? Calculate this in three ways as follows :
(a) As frame;
(b) the frame whose origin is B and whose axes remain parallel to an inertial set of axes;
(c) the rotating frame centred at A and rotates around A with the same frequency as
that of B.
Appendix B
QUIZ on The Quantum Theory
The followings are miscellaneous problems taken from sources listed in the bibliography.
1. Consider the following metals : lithium, beryllium, and mercury, with their own work
functions 2.3 eV, 3.9 eV, and 4.5 eV, respectively. One of these materials is consecutively
placed as a metallic target on a vacuum tube of the photo-electric current experimental
apparatus. If light with a wavelength of 300 nm is incident upon the target, determine
which metals show the photo-electric eect, and calculate the maximum kinetic energy
of the photo-electron in each case.
2. How does the Compton eect dier from the photo-electric eect ? Consider a photon of
frequency collides with a stationary electron of rest mass m. Show that it is impossible
for the photon to impart all its energy to the electron.
3. The binding energy of an electron is the minimum energy required to remove the electron
from its ground state to a large distance from the nucleus.
(a) What is the binding energy for the hydrogen atom ?
(b) What is the radius of the rst Bohr orbit and the binding energy for He
+
?
(c) What is the radius of the rst Bohr orbit and the binding energy for Li
2+
?
Note that a singly ionized helium He
+
and a doubly ionized lithium Li
2+
behave as
hydrogen-like atoms in that such ionized elements consist of a positively charged nucleus
and a single, negatively charged bound electron.
4. Use the Bohr quantisation rules to calculate the energy levels for a harmonic oscillator,
for which the energy is p
2
/2m + m
2
r
2
/2 directly given by the driving force m
2
r.
Restrict yourself to circular orbits. What is the analog of the Rydberg formula ? Show
that the corresponding principle is satised for all values of the principal quantum
number n used in quantizing the angular momentum.
71
72 B. QUIZ on The Quantum Theory
5. An electron has a de Broglie wavelength equal to the diameter of the hydrogen atom.
What is the kinetic energy of the electron ? How does this kinetic energy compare with
the ground-state energy of the hydrogen atom ?
6. For an electron to be conned to a nucleus, its de Broglie wavelength would have to be
less than 10
14
m.
(a) What would be the kinetic energy of an electron conned to this region ?
(b) On the basis of this result, would you expect to nd an electron in a nucleus ?
7. Atomic nuclei, typically of size 10
14
m, frequently emit electrons, with energies of
about 1 MeV or less. Use the uncertainty principle to show that hypothetical electrons
trapped electrons within a conned region having an energy of order 20 MeV could
not be contained in the nucleus before the decay.
8. Use the Heisenberg uncertainty principle to estimate the ground state energy of a one
dimensional harmonic oscillator, where the total energy of the oscillator is given by
E =
p
2
2m
+
1
2
m
2
x
2
9. A
o
meson is an unstable particle produced in high energy particle collisions. It has a
mass-energy equivalent of about 135 MeV, and it exists for an average lifetime of only
8.7 10
17
second before decaying into two rays. Using the uncertainty principle,
estimate the fractional uncertainty m/m in its mass determination.
10. One of the cornerstones of quantum mechanics is that bound particles cannot be in
stationary states even at zero absolute temperature. A bound particle is one that
is conned in some nite region of space, as is an atom in a solid. There is a non-zero
lower limit on the kinetic energy of such a particle. Suppose a particle is conned in
one dimension to a region of width L. Obtain an approximate formula for its minimum
kinetic energy.
Bibliography
Beiser, A. 1988 Perspective of Modern Physics. London, UK: McGraw-Hill.
Gasiorowicz, S. 1996 Quantum Physics. New York, US: John Wiley and Sons.
Harris, R. 2007 Modern Physics. California, US: Pearson Addison-Wesley.
Liboff, R. L. 1980 Introductory Quantum Mechanics. Reading, US: Addison-Wesley.
McMahon, D. 2006 Relativity Demystied. New York, US: McGraw-Hill.
Morin, D. 2003 Introductory Classical Mechanics. New Jersey, US: Prentice Hall.
Serway, R. A., Moses, C. J. & Moyer, C. A. 2005 Modern Physics. California, US:
Thomson Learning Inc.
Tipler, P. A. 1999 Physics for Scientists and Engineers. New York, US: W. H. Freemann.
73

Anda mungkin juga menyukai